Skip to main content

Full text of "Helicopters calculation and design. Volume I - Aerodynamics"

See other formats


I 



NASA TECHNICAL 

TRANS L ATION 



On 



»•» 



HELICOPTERS 

CALCULATION AND DESIGN 
Volume I. Aerodynamics 

by M. L. MiP et al. 

"Mashinostroyeniye" Publishing House 
Moscow, 1966 




NASA 
TT 

v.l 
c.l 



11 NASA TT F-494 









NATIONAL AERONAUTICS AND SPACE ADMINISTRATION • WASHINGTON, D. C. • SEPTEMBER 1967 



TECH LIBRARY KAFB, NM 

DDba^71 
NASA TT F-494 



HELICOPTERS 

CALCULATION AND DESIGN 
VoL I. Aerodynamics 



By M. L. Mil', A. V. Nekrasov, A. S. Braverman, 
L. N. Grodko, and M. A. Leykand 



Translation of "Vertolety. Raschet i proyektirovaniye . 1. Aerodinamika." 
Izdatel'stvo Mashinostroyeniye, Moscow, 1966. 



NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 



For sale by the Clearinghouse for Federal Scientific and Technical Information 
Springfield, Virginia 22151 - CFSTI price $3.00 



MNOTATION /g' 

The work "Helicopters (Calculation and Design)" is published in three 
volumes . 

Vol.1 - Aerodynamics; 

Vol.11 - Vibrations and Dynamic Strength; 

Vol. Ill - Design. 

The first volume is devoted to ways of developing helicopters, the basic 
principles of their design, and the position occipied by helicopters among other 
means of aviation not requiring airfields. Various theories of rotors and cor- 
responding methods of determining their aerodynamic characteristics are pre- 
sented: the classical theory of a rotor with hinged blades in the general case 
of curvilinear flight of the helicopter; the momentum theory of an ideal rotor 
and its application to the energy method of calculation; the classical theory 
when using methods of numerical quadrature; the vortex theory and methods of 
experimental determination of rotor performance in flight tests and in wind 
tunnels. Various methods of aerodynamic calculation of a helicopter and the 
theory of blade flutter are presented in detail. This volume gives an account 
of methods of calculating flutter in hovering and in forward flight. Particular 
attention is devoted to consideration of friction in the axial hinges of the hub 
and to the transfer of blade vibrations through the automatic pitch control 
mechanism. Experimental investigations of flutter are described. 

The book is intended for engineers of design offices, scientific workers, 
graduate students, and teachers of higher institutes of learning. It might be 
useful to engineers of helicopter manufacturers and to students for furthering 
their knowledge of the aerodynamics and mechanical strength of helicopters. 
Mary sections of the book will be a useful tool also to flight and technical 
staffs of helicopter flight units. 



■«• 



^ Numbers in the margin indicate pagination in the original foreign text. 
11 



PREFACE 1^ 

The present book generalizes the experience of the scientific work and 
practical design activity of engineers of one of the Soviet teams working on the 
development of helicopters. 

Twenty years ago, when the team had just set out on their work, everything 
in this field seemed to have been already long discovered and in-vented. 

Those to whom belongs credit for the original ideas and designs of rotary- 
wing aircraft - Leonardo da Vinci, M.V.Lomonosov, M .Ye . Zhtikovskiy (Joukowski), 
B.N.Yur'yev, and others - had long ago proposed almost all of the existing de- 
signs of helicopters. Designers, scientists, and inventors in -various countries 
birilt dozens of helicopter models which successfully rose into the air. However, 
not one of these rotocraft was suitable for practical use, large-scale produc- 
tion, or reg\£Lar ser-vice. 

A very difficult problem that required considerable and tedious work re- 
mained ixnsol-ved, namely, the problem of de-veloping helicopters which would find 
practical use in everyday life. 

To solve this problem we had at our disposal an inportant scientific basis 
in the form of classical works, the studies of the Central Aero-Ifydrodynamic 
Institute (TsAGi), and of foreign scientists. However, testing of each new air- 
craft confronted design engineers with new acute problems and forced them to 
work out many theoretical problems to find the proper method of sol-ving specific 
design problems. 

This volume discusses the basic problems of the theory, calculation, and 
design of helicopters worked out by the team and representing the vital interests 
of its design acti-vity. 

The fact that some of the authors had occasion to participate in applying 
the classical rotor theory to the calculation and design of the first autogiros, 
in the original experimental work on models and on full-scale rotors in wind 
tunnels, in de-veloping methods of aerodynamic calculation of helicopters, and 
then - for more than 15 years - in designing an entire family of helicopters of 
the same configuration in all weight classes, offers an opportunity to elucidate 
the basic problems of the theory and calculation of helicopters that have been /4 
checked out by practice. 

As early as 194S there was not a single helicopter in service in our 
country. Now thousands of such machines created by various design teams assist 
people in many areas of their life and activity. 

Engineers and designers working on the design or construction of heli- 
copters, pilots and technicians, students of air academies who are studying or 
are interested in helicopters will find useful information in this book. 

iii 



Engineering, especially aircraft engineering, is rapidly becoming obsolete. 
However, it is hoped that the general methods of approach to the development of 
a new type of aircraft, as presented in this book, will outlive today's heli- 
copter models. 

M.Mil' 



Chapter I of Vol.1, Sections 1 and 2 of Chapter II, and Section 2 of Chap- 
ter III were written by M.L.Mil'; Chapter I? and Section 5 of Chapter II were 
written by A.V.Nekrasov; the remaining Sections of Chapters II and III and also 
Subsections 19-28 of Section 2 of Chapter II were written by A.S.Braverman. 

In preparing the manuscript, the authors were assisted by engineers F.L. 
Zarzhevskaya, R.L.Kreyer, and L.G.Rudnitskiy. 

Reviewer R.A.Mikheyev made many valuable comments. 

The authors express their sincere gratitude to these coworkers. 



XV 



TABLE OF CONTENTS 



Page 



Preface ill 

Notations xLii 

CHAPTER I EVOLUTION HISTORY OF HELICOPTERS AND BASIC 

DESIGN PRINCIPLES 1 

Section 1. Evolution of the Helicopter Industry 1 

1. Development of Helicopters in Size 3 

2. Qualitative Development of Helicopters 8 

3. Special-Purpose Helicopters 13 

4. Compound Helicopters with Additional 

Engines - Rotocraft 15 

Section 2. The Helicopter Compared to Vertical Takeoff 
and Landing and Short Takeoff and Landing 

Aircraft l6 

1. Tactical and Technical Requirements for VTOL 

and STOL Military Transport Aircraft of the West .. 17 

2. Means for Increasing the Flying Range of 

Helicopters 21 

3. Helicopter with Takeoff Run 23 

4. Takeoff Distance of Helicopter 25 

5. Criterion for Estimating the Economy of 

Various Transport Aircraft 27 

6. Possibilities of Increase in Maximum Fljring 

Speed 31 

Section 3« Basic Principles of Design 33 

1. Selection of Engine Horsepower and Rotor Span ... 33 

2. Analysis of Multirotor Configurations 39 

CHAPTER II ROTOR AERODYNAMICS 45 

Section 1. Development of Rotor Theory and Methods of 

Experimental Determination of its Characteristics ... 45 

1. Classification of Rotor Theories 54 

2. Development of Experimental Methods 54 

Section 2. Classical Theory of a Rotor with Hinged Blade 

Attachment; General Case; Curvilinear Motion 56 

Rotor Theory in Curvilinear Motion 57 

1. Coordinate System and Physical Scheme 

of the Phenomenon • 57 

2. Inertia Forces Acting on the Blade 59 

3. Aerodynamic Forces Acting on the Blade 65 

4. Equation of Moments Relative to Flapping 

Hinge .« 66 

5. Physical Meaning of the Obtained Result 70 

6 . Equation of Torque 72 



■■■■■■ ■■■■■■ 



Page 

7. Rotor Thrust and Angle of Attack 74 

8. Lateral Force 75 

9 . Longitudinal Force 77 

10. Consideration of the Change in the Law of 
Induced Velocity Distribution during 

Curvilinear Motion 7S 

Analysis of Obtained Results 83 

11. Blade Flapping 83 

12. Effect of Curvilinear Motion at Autorotation 

of the Rotor 86 

13. Behavior of the Resultant of Aerodynamic 

Forces in Curvilinear Helicopter Motion 88 

Effect of Rotor Parameters and Hub Design on Flapping 

and Damping of the Rotor 91 

14. Rotor with a Profile Having a Variable 

Center of Pressure 91 

15. Effect of Blade Centering 92 

16. Rotor with Flapping Compensator 94 

Rotor Flapping in Curvilinear Motion of the Rotor Axis 

at Variable Angular Velocity 96 

17. Uniformly Accelerated Rotation of the 

Rotor Axis 96 

18. Harmonic Oscillation of the Rotor Axis 100 

Characteristics of Rotor Aerodynamics Determined by 

Hinged Blade Attachment 102 

19 . Physical Meaning of Blade Flapping 103 

20. Redistribution of Aerodynamic Forces over 

the Rotor Disk due to Flapping 104 

21. Approximate Derivation of Formulas for 

Flapping Coefficients 107 

22. Effect of Nonuniformity of the Induced 

Velocity Field on the Flapping Motion 109 

Method of Calculating the Aerodynamic Characteristics of 

a Rotor for Azimuthal Variation of Blade Pitch 114 

23. Equivalent Rotor Theory 114 

24. Derivation of Formulas for a Rotor with 
Flapping Hinges as for a Rotor without Hinges. 
Conditions of Equivalence of Hinged and 

Rigid Rotors 123 

25. General Expressions for Determining the Com- 
ponents of Blade Pitch Change cfb , cpi , and cpi 132 

26. Determination of Flapping Coefficients of 

Rotor with Flapping Compensator 138 

27. Determination of the_ Components of Blade 
Pitch Change cpi and 91 after Deflection of 

the Automatic Pitch Control 140 

28. Sequence of Aerodynamic Calculation of a 

Rotor with Variable Pitch •...*. 144 

Section 3 . Momentum Theory of Rotor 14^ 

1. Theory of an Ideal Helicopter Rotor 147 



VI 



Page 

2. Derivation of the Expression for the Torque 
Coefficient of a Real Rotor 156 

3. Rotor Profile Losses ...« 160 

4. Certain Considerations in Selecting Blade 

Shape and Profile l64 

5. Approximate Determination of Rotor Profile 

Losses *. 169 

6. Effect of Air Compressibility of Rotor 

Profile Losses 170 

7. Induced Losses of a Real Rotor 178 

8. Determination of Angle of Attack and 

Pitch of Rotor 183 

Section 4- Classical Rotor Theory. Method of Numerical 

Integration • 184 

1. Formulas for Calculating Forces and Moments 

of a Rotor 185 

2. Method of Calculation 193 

3. Aerodynamic Characteristics of Profiles 

for Rotor Blades 195 

4. Distribution of Aerodynamic Forces over 

the Rotor Disk 200 

5 . Aerodynamic Characteristics of Rotor 206 

6. Aerodynamic Characteristics of Rotor in 

Autorotation Regime 209 

7. Limit of Permissible Helicopter Flight 

Regimes (Flow Separation Limit) 212 

8. Distribution of Profile Losses over Rotor Disk. 
Dependence of Profile Losses on Aerodynamic 
Characteristics of Blade Profiles 218 

Section 5. Vortex Theory of Rotor 222 

1. Problems in Vortex Theory 222 

2. Theoretical Schemes for the Vortex Theory of a 

Rotor with a Finite Number of Blades 224 

3. Form of Free Vortices 226 

4. Determination of the Induced Velocities by 

the Biot-Savart Formula 227 

5. Use of the Biot-Savart Formula in Developing 

the Vortex Theory of a Rotor 228 

6. Axial Component of Induced Velocity from 

Bound Vortices 230 

7. Axial Component of Induced Velocity from 

Spiral (Longitudinal) Vortices 230 

8. AicLal Component of Induced Velocity from 

Radial (Transverse) Vortices 232 

9. Integrodifferential Equation of the Vortex 

Rotor Theory 232 

10. Constancy of Circulation of Trailing Vortices 
along Straight Lines Parallel to the Axis of 
the Inclined Vortex Cylinder and Possible 
Simplifications 234 



vxx 



Page 

11. Characteristics of Using the Lifting-Line 

Scheme and Scheme of a Vortex Lifting Siirf ace .... 236 

12. Division of Vortices into Types Close to 
and Remote from the Blade; Use of "Steady- 
Flow Hypothesis" 237 

13. Instantaneous and Mean Induced Velocities 
and Generation of Variable Aerodynamic 

Loads on the Blade 238 

14. Characteristics of the Extrinsic Induced 

Velocity Field 238 

15. Vortex Theory of a Rotor mth an Infinite 

Number of Blades 239 

Vortex Theory of Wang Shi-Tsun 240 

16 . Rotor Scheme 240 

17. Determination of Induced Velocities 241 

18. Calculation Formulas for Induced Velocity 
Determination 241 

19. Application and Evaluation of the Possibilities 

of the Wang Shi-Tsun Vortex Theory 243 

Vortex Theory of V.E.Baskin 244 

20. Scheme of Rotor Flow 245 

21. Determination of Induced Velocities from 

the Dipole Column 246 

22. Fluid Flow Induced by a Disk Covered with 

Dipoles 247 

23 . Boundary Conditions 249 

24. Transformation of Eq.(5.67) to the Rotor Axes; 
Use of the Theorem of Addition of Cylindrical 
Functions 249 

25. Determination of the Total Velocity Potential 

from the Entire Dipole Column 250 

26. Determination of Induced Velocities 252 

Section 6. Experimental Determination of Aerodynamic 

Characteristics of a Rotor 253 

1. Flight Tests for Determining the Aerodynamic 
Characteristics of a Helicopter 254 

2. Wind- Tunnel Tests for Determining the 

Aerodynamic Characteristics of a Rotor 257 

Methods of Converting the Aerodynamic Characteristics 

of a Rotor 26l 

3. Conversion of Aerodynamic Characteristics to 

a Different Rotor Solidity Ratio 26l 

4. Conversion of Aerodynamic Characteristics on 
Variation in Minimum Profile Drag Coefficient 

of the Blade Sections Cxpg • 265 

5. Conversion of Aerodynamic Characteristics on 
Variation in the Peripheral Speed of the 

Rotor (Mo Numbers) 266 

6. Conversion of Angle of Attack and Rotor Pitch 



vxxi 



Page 

on Variation in Inclination of the Automatic 
Pitch Control, Flapping Compensator, and Mass 

Characteristic of the Blade 26? 

7. Examples of Using the Conversion Formulas 268 

Section ?• Performance and Propulsive Efficiency Coeffi- 
cient of a Rotor 270 

1. Performance and Efficiency of Rotor 

Proposed by K .Khokhenemzer 271 

2. Determination of Performance and Propulsive 

Efficiency of a Rotor 273 

3. Performance and Efficiency of a Rotor, 

Obtained from Experimental Data ........*.. 277 

4. Performance and Efficiency of a Rotor, 

Obtained from Calculated Graphs 279 

5. Conversion of Performance and Efficiency on 

Variations in Rotor Parameters 282 

6. General Comments on Rotor Efficiency and 

Performance 283 

Section 8. Calculation of Rotor Characteristics in 

Hovering and Vertical Ascent (Momentum Theory 

of Propellers) 284 

1. Brief Review of the Momentum Theory of 

Propellers 284 

2. Results of Calculating the Characteristics 

of a Rotor 286 

3. Approximate Method of Determining the 

Dependence of mt on t 292 

4. Conversion of Aerodynamic Characteristics on 

Variation in the Rotor Solidity Ratio 295 

5. Determination of Optimal Aerodynamic Parameters 
of a Rotor with Consideration of the Dependence 

of Characteristics on Mo 296 

CHAPTER III AERODYNAMIC DESIGN OF A HELICOPTER 301 

Section 1. Basic Equations for Aerodynamic Design of 

a Helicopter 301 

1. Aerodynamic Design Principle of a Helicopter .... 301 

2. Equation of Motion of a Helicopter 301 

3. Various Methods of Determining Aerodynamic 
Rotor Characteristics and Methods of 

Aerodynamic Design 303 

4. Calculation of Composite and Multirotor 

Craft 304 

5. Induction Coefficients of Two-Rotor Helicopters 

and Helicopters with a Wing 308 

Section 2. Aerodynamic Helicopter Design by the 

Mil'-Yaroshenko Method 315 

1. Equations of Motion and Design Principles 315 

2. Determination of Aerodynamic Rotor Characteristics . 318 



IX 



Page 



3. Calculation of Flight Data 320 

4. Limits of Applicability of the Method 322 

Section 3. General Method of Aerodynamic Design for 

Rotary Wing Aircraft 323 

1. Construction of A-uxiliary Graphs for Helicopter 
Performance Data 324 

2. Determination of Helicopter Performance Data .... 331 

3. Graphs for Determining Optimum Helicopter 

Aerodynamic Parameters 342 

Section 4. Aerodynamic Design of a Helicopter Based on 

Concepts of Rotor Performance and Efficiency 346 

1. Helicopter Performance 347 

2. Performance of Multirotor and Composite 

Helicopters 348 

3. Determination of Helicopter Flight Data 357 

4. Calculation of a Helicopter -with a Tractor 

Propeller 363 

5. Comparison of Helicopter and Airplane 364 

6. Power of Front and Tail Rotors in a Helicopter 

of Fore-and-Aft Configuration 366 

7. Retraction of Landing Gear on Helicopters 368 

Section 5. Aerodynamic Calculation of a Helicopter by 

the Power Method 369 

1. Determination of Required Power in Horizontal 
Helicopter Flight 370 

2. Determination of Helicopter Performance Data .... 375 

3. Relation between Np^, Ni„4, and Npaj. during 

Horizontal Flight of a Single-Rotor Helicopter , . . . 376 

CHAPTER IV ROTOR FLUTTER '. . . . 379 

Section 1. Basic Assumptions and Characteristics of an 

Approach to Flutter Calculation 380 

1. Bending and Torsional Vibrations of the Blade. 

Possible Cases of Stability Loss 380 

2. Effect of Blade Attachment to Hub and the 
Possibility of Theoretical Investigation of 

Flutter of an Isolated Blade 381 

3. Different Types of Flutter Differing with 
Respect to Blade Vibration. Flapping and 

Bending Flutter 381 

4. Characteristics of the Torsional Vibration 
Modes of a Blade and Possible Correlated 

Assumptions 382 

5. Assumptions on Blade Oscillations in the Plane 

of Rotation 383 

6. Determination of Aerodynamic Forces Acting on 

a Vibrating Profile 384 

Section 2. Flapping Flutter of an Isolated Blade with 

Axial Flow past the Rotor 386 



Page 

1. Blade Model 386 

2. Derivation of Differential Equations 

of Flutter 3S7 

3. Particular Solution of the Differential 

Equation 391 

4. Differential Equation of Disturbed Motion 391 

5. Notation of Differential Equations in 

Matrix Form 392 

6. Solution of Differential Equations of 

Blade Vibrations 392 

7. Determination of the Critical Flutter Rpm 395 

8 . Blade Divergence 396 

9. Parameters Characterizing Blade Balance 

(Effective Blade Balance) 396 

10. Dependence of Critical Flutter Rpm on Blade 
Balancing and Values of the Flapping Com- 
pensator Coefficient 398 

11. Blade Arrangement 399 

12. Effect of Control Rigidity 400 

13 . Conditions for Ab sence of Flutter 400 

14. Mechanism of Generation of Forces Exciting 

Flutter 401 

Section 3- Consideration of Friction Forces during Flutter .... 406 

1. Character of the Effect of Friction Forces 

during Flutter 406 

2. Linearization of Friction Forces 40? 

3. Determination of Flutter Speed with 

Consideration of Friction 408 

4. Effect of Forced Motion in the Feathering 

Hinge 409 

Section 4- Rotor Flutter with Consideration of Coupling 
of Blade Vibrations through the Automatic 
Pitch Control 414 

1. Forms of Rotor Flutter Observed in 

Helicopter Experiments 414 

2. Analytical Expression for Cyclic Modes of 

Rotor Vibration 414 

3. Cyclic Vibration Modes in Specific Cases 

and Control Loads 416 

4. Differential Equations of Rotor Flutter with 
Consideration of Coupling of Blade Vibrations 

through the Automatic Pitch Control 418 

5. Transformation of Eqs.(4.18) in Particular 
Cases where Cyclic Modes are the Solution of 

the Differential Equations of Rotor Flutter .... 421 

6. Rotor Flutter in the Presence of Different 

Rigidity of Longitudinal and Lateral Controls ... 422 
Section 5- Flapping Flutter of a Rotor in Forward Flight 424 

1. Preliminary Statements 424 

2. Differential Equations of Blade Oscillations 

in Forward Flight 424 



XI 



3. Solution of Differential Equations 

4. Determination of Critical Flutter Rpm 
"without Consideration of Harmonic Components 
of Blade Motion 

5. Effect of Flying Speed on Critical 

Flutter Rpm 

Section 6. Calculation of Flutter vd_th Consideration of 

Bending and Torsion of the Blade 

1. Bending and Torsion of Blade during Flutter 

2. Determination of the Torque from Bending 
Forces on the Blade 

3. Differential Equations of Binary Vibration 

4. Solution of Differential Equations 

5. Calculation of Flutter with Consideration 

of Three Degrees of Freedom 

6. Calculation of Flutter with Three Degrees 

of Freedom Disregarding Blade Torsion 

7. Calculation Results 

8 . Bending Flutter 

9. Approximate Method of Determining the Mode 

of Bending Vibrations in Flutter 

Section ?. General Method of Calculation of Flutter and 
Bending Stresses in the Rotor Blade during 
Flight 

1. Calculation Method and its Possibilities ... 

2. Basic Assumptions and Suggestions 

3 . Differential Equations 

4. Boundary Conditions of the Problem 

5. Determination of Equivalent Rigidity of the 
Control System , 

6 . Determination of Aerodynamic Forces , 

7. Method of Solving the Differential Equations 

8. Transformation of Partial Differential 
Equations into Ordinary Differential 
Equations , 

9. Determination of the Magnitude of the Moment 
of Friction in the Feathering Hinge of the Hub 

10. Sequence of Performdng the Calculation 

Section 8. Experimental Investigations of Flutter 

1. Ground Tests for Flutter 

2. Flutter Tests in Flight 

3. Comparison of Calculation and Experiment 
under Conditions of Axial Flow past the Rotor 

4. Comparison of Calculation and Experiment in 
Flight 

5 . Check for Flutter 

6. Experimental Determination of Control System 
Rigidity 

7. Experiments on Djmamically Similar Models ... 
References 



Page 
426 

427 
428 

429 
429 

430 
432 
434 

436 

441 
446 
447 

450 



452 
452 
454 
454 
455 

456 
457 
460 



462 

466 
467 
470 
470 
475 

478 

480 
480 

481 
483 
487 



Xll 



NOTATIONS l^ 

Ae rodynajnic Characteristics 

a = angle of attack of rotor; 
ffp = angle of attack of blade section; 
Qfo = angle of zero lift of blade profile; 
Aa = downwash angle of flow; 

^^^ = Cy = tangent of angle of slope of the lift curve 

■with respect to angle of attack of the profile; 



da 



= tan"'' — ^ = inflow angle in blade section; 

r = circulation in blade section; 
M = Mach number of blade section; 
Mq = average Mach number with respect to azimuth, 

in tip section of blade (Mq = -—■ 



Mfi = flight Mach number (Mfi= -|- 

Q 

T, t = — I- = thrust and coefficient of thrust of rotor 

/ _ T ^ 
l^ ~ i PctttR2((juR)^ J ' 

Q 

H, h = —^ = longitudinal force and coefficient of longi- 
tudinal force of rotor ( h = 



^ paTTR^((juR)^ 

Cs 

S, s = = lateral force and coefficient of lateral force 

a 

of rotor ( s = 



i parrR^((i}R)= 



m 



M^, mt = - = torque and torque coefficient of rotor 

Mt 
m. = 



•J- parrR^CouR)^ 



N = power of motor (of rotor in Chapt.II); 
lift and lift coeffic: 
(t ^ V 

V ^ i nrrTTR2C.,R;i2 J ' 



Cv 

Y, ty = —-I— = lift and lift coefficient of rotor 



parrR^CcuR)^ 



xiix 



Velocities 



X, tx = — ^ = propulsive force and coefficient of pro- ^ 

( X \ 
pulsive force of rotor it, = — — — — =); 

Cy» Cxp = coefficients of lift and profile drag of 

blade section (airfoil) referred to dynamic 
pressvire ^ pV^; 
B = coefficient of tip losses; 



cbR = angular velocity; 



V, fv = -^ = path velocity of helicopter flight; 

\ (bR / 



\ (bR 

^x» ^y» ^z ~ horizontal, vertical, and lateral conponents 
of flight velocity; 

V, ( V = ~ ^ - ■ ) = induced velocity; 
\ luR / 

U, ,^U = ^^1 = relative velocity of flow past a blade 

V U)R / element; 

Ux, Uy, (Ux = — ^, Uy = — 5L_| = horizontal and vertical conponents of rela- 
^ '"-^ "^^ ' tive velocity of flow past a blade element; 

\ = coefficient of flow; 

p, = characteristic (coefficient) of rotor per- 
formance . 

Geometric Characteristics 

D = diameter of rotor; 

R = radius of rotor; 

F = disk area; ,_ . 

r = radius of rotor blade section f r = -^)> 

b = blade chord ', b = -— -, b = -,- \ 

Zi, = number of blades; 
Zr = number of rotors; 

a = ^^^o-"^ = load factor of rotor; 

ttR 

^.h* ^v.h - distance from axis of rotation of rotor to 
the horizontal (flapping) and vertical 
(drag) hinges, respectively; 
c = thickness of profile section, c = -^; 
P = flapping angle of blade; 

xiv 



a„, b„ = coefficients of flapping; 

9, 9o = blade angle (pitch); angle between chord of blade 
profile and plane of rotation; 
H, T) = angles of deflection of automatic pitch control mecha- 
nism; H -with index = mutual influence coefficient of 
lifting elements; 
00 = blade angle at r = 0.7 for P = k = Tl = 0; 
3i> 9s = cQDponents of change of blade angle relative to the 

plane of rotation, due to deflection of the automatic 
pitch control mechanism; 
V = change of blade angle due to elastic deformation of 
blade. 



XV 



HELICOPTERS; CALCULATION AND DESIGN. VOL.1: AERODINAMICS /2 

M.L.Mil', Editor 

ABSTRACT. A review of the historical development of Russian 
and Western helicopters, in size and lift capacity, for civil 
and military purposes is followed by detailed discussions on 
rotor aerodynamics for various angles of attack, blade setting, 
flapping angle, center-of-pressiire position, blade vibration 
(natioral, forced, harmonic, etc.), and other rotor parameters 
in their influence on rotor rpm and craft stability. Formulas 
are given for the forces and moments of rotor danping in hover- 
ing and forward flight; for the redistribution of aerodynamic 
forces over the rotor disk due to flapping; for cyclic pitch 
change of rotors with variable and constant pitch. The theory 
of an ideal helicopter is developed on the basis of optimum 
blade profile, prevention of rotor profile losses, and proper 
balancing. Flutter in hovering and forward flight is calcu- 
lated, with eirphasis on friction in the axial hub hinges and 
transfer of vibrations through the automatic pitch control. 

CHAPTER I 

EVOLUTIONAL HISTORY OF HELICOPTERS AND BASIC 
DESIGN PRINCIPLES 
(Selection of Parameters and Configuration) 

Section 1. Evolution of the Helicopter Industry 

Designing is always directed toward the futiire. However, for a better 
picture of the potentialities of the future development of helicopters it is 
useful to attenpt to understand the basic trends of their evolution from past 
experience. Natiu-ally, we are not interested here in the prehistory of heli- 
copter construction, which we will only briefly mention, but in its history from 
the time when the helicopter as a new type of aircraft became usefiil for practi- 
cal application. 

The writings of Leonardo da Vinci going back to 1483 contained the first 
mention of an apparatus with a vertical rotor, a helicopter. The first stage of 
evolution ranges from the model of a helicopter developed by M.V.Lomonosov in 
1754 through a long series of designs, models, and even full-scale apparatus 
which were not destined to rise into the air, to the construction of the world's 
first helicopter which, in 1907, was able to become airborne. This four-rotor 
helicopter was constructed by the French designers Breguet and Riche. In 1923, 
a passenger became airborne for the first time in the USA in a helicopter de- 
signed by de Bothezat. The first world altitude record of a helicopter of 18 m 
was set in 1930 on the Italian coaxial helicopter by d'Ascanio. 



In Russia, a single-rotor helicopter was built in I9II, on the basis of 
the scientific research by M.Ye.Zhukovskiy devxated to helicopter rotors, by a 
groigj of his students headed by B.N.Yur'yev. The configurations of this machine 
represent the basic scheme of the single-rotor helicopters used widely at 
present. B.N.Yur^yev was able to resume this work only in 1925* In 1932, a 
grotf) of engineers headed ty A.M.Cheremukhin constructed the helicopter TsAGI 
1-EA (Fig. 1.1) which reached an altitude of 6OO m and stayed in the air for 
18 min, which - for that time - was an outstanding achievement. It suffices to 
say that the official altitude record established three years later on Breguet^s 
new coaxial helicopter was only 180 m. 

At this time there was a pause in the development of helicopters. A new 
branch of rotary-wing aircraft came to the forefront, known as autogiros. The 
idea of the autogiro, as an aircraft with a rotary wing (freely rotating air- 
foil) never losing speed, occurred to the young Spanish engineer Juan de la 
Cierva in the 1920s. At that time, conventional aircraft whose development had 
been vigorous during the years of World War I and which, by then, carried /8 
armament and thus had greater wing loading were troubled by a new problem of 
spin, i.e., stalling. It appeared sinpler to develop a safe and sufficiently 
perfected autogiro than to build a helicopter. The rotor, freely rotating due 
to the relative flow, eliminated the need for conplex reduction gearing and 
transmissions. The hinged attachment of the rotor blades to the hub used on 
autogiros gave far greater strength to the blades and higher stability to the 
autogiro. Finally, engine failure ceased to be a threat, as had been the case 
in the first helicopters; the autogiro, with autorotating blades, had no diffi- 
ciolty in landing at low speed. 



Fig. 1.1 Helicopter TsAGI 1-EA. 

Cierva, working in England, created several autogiro designs, the best 
known of which was the C-30 autogiro which was produced as a pilot series. 
Autogiros were also built in the USA by the Pitcairn and Kellett Companies and 
in the Soviet Union at TsAGI by the designers I.P.Bratukhin, V.A.Kuznetsov, 
N.I.Kamov, .N.K.Skrzhinskiy, M.L.Mil', and others. 

The flying speed of Soviet autogiros in 1937 reached 260 km/hr. The A-7 
autogiros designed by N.I.Kamov were used at the front during the first year of 
World War II. 



The great lift capacity of the rotating rotor gave the autogiro a short 
ground run. Even though, a mechanical drive from the engine, for spinning the 
rotor before takeoff, was used in this design to further shorten the takeoff 
run. In the design of the British C-40 autogiro the rotor was given a spin-Tjp 
before fUght to an rpm such that, at the instant of disengagement from the 
engine - which, in forward flight, rotated the propeller - the machine, due to 
the marked increase in pitch, took off without a run, rising vertically into the 
air. 

Only one step remained for the development of a true helicopter. And this 
step, as is always the case in technology, was made almost simultaneously in 
various countries. This was the beginning of the present development stage of 
helicopters . It was started by flights • of the FW-61 helicopter designed by 
Professor Focke in Germany (1937), the VS-300 helicopter designed by Sikorsky 
in the USA (1939), and the "Omega" helicopter designed by I.P.Bratukhin in the 
USSR (1940). All three of these helicopters used a hinged rotor capable of 
autorotation, which had already become standard for autogiros . 

World War II somewhat delayed the development of helicopters. They were 
still unsuitable for practical use, and the ways and means for experimental 
studies were limited. After the end of the war (1946 and 1947), large numbers 
of designers and inventors invaded this new and promising area of development of 
aviation engineering. Within a short time, literally dozens of new helicopter /2. 
designs were created. This was a contest of the most diverse schemes and con- 
figurations, generally of the single- or two-seater type and used mainly for 
experimental purposes • Military agencies were the only users of this expensive 
and conplex equipment. The first helicopters in various countries were used as 
liaison and reconnaissance military aircraft. 

In the development of helicopters, just as in many other areas of tech- 
nology, one can clearly distinguish two trends of development: the quantitative 
trend concerned with size of the machine and the almost simultaneous qualitative 
trend concerned with inprovement of the craft within a certain size or weight 
class . The former trend represents development with respect to lift capacity 
and the second with respect to inprovement of the tactical or economic features 
of helicopters. 

1. Development of Helicopters in Size 

A study of foreign helicopters indicates that the use of helicopters for 
landing Marines from ships was the determining factor in the further development 
of military helicopters as troop carriers. The American landing of troops in 
S-55 helicopters at Inchon during the Korean War (1951) was a typical exanple 
of this trend. 

The size range of the assault helicopters was predicated on bulk and weight 
of ground transportation means used by the troops and to be dropped by air. It 
is a known fact that conventional weapons - mainly artillery - transported by 
prime movers are close in weight to the weight of the prime movers themselves. 
Thus, the lift capacity of the first transport helicopters in annies of other 
countries was 1200 - I6OO kg (the weight of a light military truck used as 



prime mover together with the re^ective weapons). Subsequently, the required 
lift capacity of helicopters was increased to 6 - 8 tons which, in accordance 
with military technique, was based on automobile carriers with a lift capacity 
of 3 - 4 tons. Still later, for exanple in projects developed by Sikorsky Air- 
craft, the lift capacity of helicopters rose to 20 - 25 tons and finally to 
36 - 40 tons. Such weights correspond to the weight of light and mediim tanks 
or of self-propelled landing craft. Whether .this development trend in size inr- 
crease will ever come to an end depends on the constantly changing military 
planning. Artillery systems are being largely replaced by missiles, for which 
reason the foreign press often mentions the need to transport missiles or missile 
systems, the prime factor in determining the size of modern helicopters. 

In the atteiifjt to single out the main trend of future helicopter develop- 
ment, after successively outlining the creation of new types of machines in the 
few designer firms that have been successful in developing experimental models 
into practical prototypes and in starting pilot series, it will be found that 
the major development was toward an increase in the lift capacity of helicopters. 



TABLE 1.1 







Characteristics 


USSP 








Mi-1 


Mi -4 


Year of production 


1948 


1952 


Lift capacity, in 
ton- force 


0.3 


1.2—1.6 


Increase over previous 
model 




4 


Flying weight, 


2.3 


7.2 


ton- force 







Heli copters 



1957 
8—12 



39—41 



USA 



S-51 S-58 S-64 DS-103 



1946 


1956 


1962 


Projects 


0.3 


1.2 


5-6 


20 






3 


4 


3 




2 


6 


17.0 


— 





40 
2 



Table 1.1 gives data characterizing the development of the lift capacity 
of single-rotor helicopters of the same configuration by two aircraft construc- 
tion departments - helicopters Mi-1 (Fig. 1.2), Mi-4 (Fig.l.3), Mi-6 (Fig. 1.4), 
Mi-10 (Fig.1.5), S-51 (Fig. 1.6), S-58 (Flg.1.7), and S-64 (Fig. 1.8). 

As we see from Table 1.1, the lift capacity increases severalfold with 
each prototype. 

However, it is easy to show that an increase in size and weight of heli- 
copters is impossible without a qualitative inprovement of the engines used /lO 
(reduction in weight per unit horsepower and increase in econony, i.e., decrease 
in fuel consunption) . 

Actually, an increase in flying weight is possible either by increasing the 
rotor span or the installed power, or both factors 



G = T=(k^ND)'l'. 



(1.1) 



The weight of the engine is proportional to the first power of its output, 
while the weight of the machine itself increases only in proportion to the 2/3 
power . 




Fig. 1.2 Mi-1 Helicopter. 



Thus, a helicopter with a larger power-to-weight ratio will have a rela- 
tively greater design weight, owing to the power plant. 

In like manner the weight of the blade and, accordingly, the weight of the 
lifting system change in proportion to the third power of the diameter, whereas 
the weight of the helicopter again changes only in proportion to the 2/3 power. 
Here also, the weight of the lifting system of a larger helicopter proves to be 
relatively greater. Thus, on increasing the size of a helicopter its load 
ratio, i.e., the ratio of useful load to flying weight, should be decreased, if 
there is no weight inprovement in engines, blade design, reduction gears, or 
transmissions. Actually, in the 1930s papers were published that demonstrated 
the uselessness of developing helicopters with a power greater than 500 tp, 
since an increase in power would not lead to an increase in usefiol load. /13 
According to technical specifications of that time, the weight of rotors, reduc- 
tion gears, and of the entire machine as a whole increased with increasing power 
more rapidly than the lift. 

However, in developing a new militaiy - and especially a new general- 
purpose - helicopter, the designer will not tolerate a lowering of the achieved 
level of load ratio. 

Thus, a "quantitative" development with respect to size is inpossible with- 
out a qualitative development; in fact, it always is concurrent with the qualita- 
tive advance of technology. 

The development of helicopters larger than the first two- or three-place 
models took place in a conparatively short time, since the unit weight of piston 
engines always decreased with an increase in power. But in 1953, after develop- 
ment of the 13-ton Sikorsky S-56 helicopter (Fig.l.lO) with two 2300- tp piston 



m 




Fig. 1.3 Mi-4 Helicopter. 




Fig. 1.4 Mi-6 Helicopter. 




Fig. 1.5 Mi- 10 Helicopter. 



m. 




Fig. 1.6 S-51 Helicopter. 




Fig. 1.7 S-58 Helicopter. 




Il.g.1.8 S-64 Helicopter. 



^takeoff 'on-force 
40 



'm ^"«i^AM«ki^B - 



engines, the size series of helicopters in the West was discontinued and only 
in the USSR was it possible, in 1957, to develop the Mi-6 helicopter with a 
flying weight of 40 tons by using turboprop engines. 

2. Qualitative Development of Helicopters 

In the middle of the 1950s, the reliability of helicopters became appreci- 
ably greater so that also their use potentialities for the national econon^r in- 
creased. This moved problems of econong^ into the foreground. 

The operating cost per hour of a helicopter plays a decisive role in 
whether to use them for geological surveys, in agricultxore, or for transporting 
passengers. Amortization, i.e., the price of a helicopter divided by its 
service life, constitutes a large portion of the cost. The service life of the 
helicopter is determined by the durability of its conponents. The problem of 
increasing the fatigue strength of blades, shafts, transmissions, rotor hubs, 
and other lonits of the helicopter became a prime problem, which helicopter de- 
signers are still studying at 
present. Today, a life of 1000 
hours is no longer a rarity for 
series-produced helicopters and 
there are no grounds to doubt its 
further increase. When using 
helicopters in transportation, 
the concepts of cost per ton-irdle. 
of the transported load and the 
cost per passenger-mile become 
decisive. This is the hourly op- 
erating cost divided by hourly 
productivity, i.e., by the product 
of the weight of the pay load and 
the cruising speed. 

Since the construction weight 
largely determines the price of a 
helicopter, the direct relation 
between econon^r and load ratio of 
the helicopter is obvious. Flying 
speed also acquires a new role. 
/m 1950 IS55 I960 I9B5 

This automatically leads to 
Fig. 1.9 Size Evolution of Helicopters. the idea of developing helicopters 

with higher economic indexes. The 
development of turboprop engines 
with an appreciably smaller unit weight than piston engines made it possible to 
produce helicopters with a larger load ratio while retaining, in each weight 
category, the rotor dimensions. 

Generally, replacement of piston engines by turboprop engines not only 
results in a decrease in relative weight of the power plant but also in some 
increase in power; produces a dual effect and also leads to an appreciable 




^^^5l3r^' 



i 



S 



increase in criiising speeds. 

In the diagram (Fig.1.9) we traced these quantitative and qualitative de- 
velopment trends of the most common helicopters produced by the three design 
engineering departments. Given are the single- rotor helicopters designed by 
SikorslQT- Aircraft (USA), the single-rotor Soviet helicopters, and the fore- 
and-aft helicopters of the Piasecki Aircraft Corporation, which subsequently 
became the Vertol Div. of Boeing. 



I2k 




Fig. 1.10 S-56 Helicopter. 

Thus, the size development trend on the basis of piston engines (soUd 
lines in Fig.l.9) was terminated as early as 1953. Then, as turboprop engines 
of the necessary size were developed over a period of five to ten years, second- 
generation helicopters appeared (points referring to these in the diagrams are 
connected with the original models by the broken line of qualitative develop- 
ment). 

Thus, the helicopters S-55, S-58, and S-56 with piston engines served /16 

as prototypes, respectively, for the turboprop machines S-6I (Fig. 1.11), S-62 
and S-65 (Fig. 1.12;. The same holds for the fore-and-aft helicopters of the 
Vertol Div. of Boeing V-107 and YBr-Hk "Chinook" (Fig.1.13). 

The Soviet turboprop helicopters Mi-2 (Fig. 1.14) and Mi-8 (Fig.1.15) also 
constitute a ftirther development of the well-known helicopters Mi-1 and Mi-4. 

The unusually long service life of helicopters is striking in conparison 
with airplanes. Almost all piston helicopters shown in the diagram (with the 
exception of the e^qjerimental helicopters XH-16 and S-56) were in production and 
service before the appearance of their second turboprop generation, and the Mi-1 
helicopter has managed to stay in production for 15 years and is approaching 
the record longevity of the Li-2 airplane. 

We can assume that the weight categories of helicopters indicated in 



|[ 



1 *^*s,^''%>^;-!:; 




m. 



Fig. 1.11 S-61 Helicopter. 




Jig. 1.12 S-65 Helicopter. 




Fig. 1.13 Chinook Helicopter. 



10 



Table 1.2 have 'beconie established hy now. 

What -will be the future development of helicopters? 



/M 



The process of developing a new generation of helicopters, on the basis of 
inp roved turboprop engines, is now being conpleted in the lightest category of 
helicopters. The lag in this weight category can be attributed to difficulties 







Fig, 1.14 Mi-2 HelLcqpter. 




Jig. 1.15 Mi-8 Helicopter. 



in developing a lighter and simultaneously more economic low-power t^lrboprop 
engine in conparison with piston engines. In the end, such an engine was de- 
veloped in the USA by the Allison Company - this was the T-63 weighing only 
174 lbs at a power of 3I5 hp and a consunption of 280 gtn/hp-hr. The award in 
the conpetition for a light three- or four-place military helicopter in the USA 
was made to the Hughes Aircraft Conpargr, which created the UH-6A helicopter 
(Pig. 1.16) weighing only 2680 lbs at an enpty weight of about 134O lbs; this is 

11 




m. 



Fig. 1.16 Hughes Helicopter UH-6A. 




Fig. 1.17 Fairchild HLller Helicopter FH-1100. 



an appreciable technical achievement which required a number of new design solu- 
tions, in particular the use of a rotor with an elastic spring retention of the 
blades instead of the conventional hinge attachment. This helicopter has a high 
load ratio {50%) combined with a high cridsing speed (213 km/hr), for a light 
machine. The Fairchild HLller FH-llOO is also in this class (Fig. 1.1?). It is 
obvious that these helicopters considerably outstrip the light Haison recon- 
naissance aircraft of World War II, both with respect to speed and lift capacity 
and, fiirthermore, have the great advantage of vertical takeoff and landing. 
Thus, the decision made in a number of countries to replace light reconnaissance 
aircraft by helicopters is not surprising. 



12 



TABLE 1.2 



Ch 



aracteri sties 



Lift capacity or 
number of places, 

Flight weight 



Light 
!.i ai son 



2 — 4 per- 
sons 

1.5—2 ton 



Light 
Multi- 
purpose 



Type of Helicopter 



Light 
Transport 



1 ton or 3 ton or 
10—12 per- 25—30 per- 



3.5 — 4 ton 



10—12 ton 



Medium 
Transport 


Heavy 

Transport 


Superheavy 
Transport 


6 — 8 ton 
20— 40 ton 


20 ton 


40 ton 



Of course, a new generation of light helicopters will also be developed in 
other countries of the world. In France, this is being done on the basis of the 
350- Ip Turbomecca-Oredon-III engine. In West Germany, the Bolkow ConparQr is 
working on such a machine. 

Thus, in speaking of the qualitative development trend of helicopters, it 
is obvious from the foregoing that each new generation of engines gives rise to 
a new generation of helicopters in all weight categories, simultaneously having 
greater econonQr and better flight performance data. This line of development 
probably has no upper limit. 

As regards the size evolution of helicopters, no machine with a lift capaci- 
ty of 20 tons (see Table 1.2) has been developed as yet. 

According to a request for proposals, announced in the USA, firms such as 
Kaman, Fairchild Killer, and Sikorsky Aircraft are working on the development of 
a helicopter with a lift capacity of 20 tons. In West Germany, the Bolkow 
Conpany is working on a helicopter with a 40-ton lift capacity. Below, we will 
review the possible ways of developing heavy and siperheavy helicopters. 

3« Speci al-Purpose Helicopters /19 

It is necessary to mention also the development of various models of 
special-purpose helicopters within the indicated weight categories. In this 
connection, let us make a brief remark on the new concept of using helicopters 
in the Army which has recently developed in the West - especially in the USA - 
namely, the creation of so-called airborne mobile troops. 

In this instance, helicopters are used in place of motorized transport for 
all types of troop movement. The Bell "Iroquois" helicopter UH-ID (Fig. 1.18) is 
particiJ.arly adapted for transporting troops by platoons (11-12 men). 

light reconnaissance three- or foua>-place armed helicopters (Hughes heli- 
copters 0H-6a), flying in front of battle formations, are also a necessity. 
Finally, regular troop-carrier helicopters of various classes, supplying the 



13 



means of gro-und fire sipport such as artillery, rockets, and tanks, take over 
the task of troop movements. 

Also used in realization of this concept are helicopters for air sipport of 
infantry, constituting a vinique type of assault helicopters. Ordinary heli- 
copters armed with radio-controlled missiles and weapons are presently used for 
this purpose. 




Fig. 1.18 Bell "Iroquois" Helicopter UH-ID. 



Such an airborne motile division is supplied from the air by airplanes and 
helicopters of the Mr Force Materiel Command. 

It is not difficult to detect behind this concept past military experience, 
wherein any new type of transportation that became accessible engendered a new 
type of troops. Beginning with cavalry, we recall the bicycle and motorcycle 
xinits of World War I, and the motorized infantry, motorized divisions, and air- 
borne troops of World War II. 

It is clear by now that this concept is finding followers in many western 
coiintries* 

Thus, the l2-place SA-330 (Fig. 1. 19) helicopter ordered by the French Army 
corresponds to the 11-place Iroquois helicopter (USA). A similar machine is 
being designed also in West Germany. 

The need to retain the class of 10- to l2-place light transport helicopters 
is confirmed also by the practical esq^erience with the l2-place Mi-Zt- helicopters 
in the national economy. It is obvious that the development of more economic 
(for airlines) 30-place Mi -8 helicopters does not interfere with the advantage 
of using the 10-place helicopters in the national economy for geology and other 
purposes. 



14 



4« Compound Helicc^ters -with Additional Engines - Rptocraft /20 

Of considerable interest was the appearance of conpound helicopters which 
use propellers for forward flight, as autogiros did earlier. Such are the 
Rotodyne Ferry designed by Hislqp and especially the rotocraft of the Soviet 
designer M.I.Kamov. 

In 1964, world records for machines of this type were set on the rotocraft 
Ka-22: speed 36O km/hr, Hft capacity 16 tons. N.I.Kamov's rotocraft again 
focused attention of the helicopter world, after 20 years, on the side-by-side 
configuration which had been successf uUy developed by Focke in Germany and by 
I.P.Bratiikhin in the USSR. This machine recalled the great advantages of the 
side-by-side configuration in flying range and lift capacity with a running 
takeoff which must be accoTinted for in a successful design. 



'**fe£ 



(.. mf''^"'"' ^'^^^ 



^•^•^Nsi, 







Fig. 1.19 SA-330 Helicopter. 



A further development of conpound helicopters with propeller is represented 
by the helicopter prototype with additional turbojet engines now being proposed 
in the West for military purposes. 

An interesting prototype of an assault helicopter is the Lockheed composite 
helicopter (Fig. 1.20;. This two-place experimental machine, in addition to the 
main 550-lnp turboshaft engine driving a four-blade rotor with elastic blade re- 
tention, uses a turbofan engine mounted on a small wing and permitting rev-up 
to 426 km/hr when briefly cut in during flight . 

The successful development of dual-flow turbofan engines, especially with 
a large bypass ratio, may lead to the development of models which, at cruising 
speed, would have a specific consimption of the order of Cr = 0.5 kg/kg ♦ hr. 
Since 

^ _ 75r, 



V 



■ C 



/?• 



it is not difficult to calculate that, in this case, the consunption per horse- 
power of an equivalent propeller engine at a propeller efficiency of 0.75 and a 
flying speed of I50 m/sec is only about 200 ^n/(lp - hr) . 

15 



If we also take into accoiuit the small weight of such a motor in conparison 
with the weight of a tTH-boprop engine, it becomes clear that the use of turbofan 
engines of this type can be economically ad-vantageous even at lower cruising /2l 
speeds and may lead to the development of conpound helicopters with an auxiliary 
thrust engine and wing for passenger transport between urban centers at cruising 
speeds of the order of 350 - 450 km/hr. At the same time, such helicopters may 
find military use as fire-stpport craft for troops. 




Fig. 1.20 Lockheed Helicopter XI^51A. 

In analyzing the ways and means of helicopter development, one cannot side- 
step the question of vertical takeoff aircraft. Will the development trend and 
use of helicopters come to an abrupt end with the appearance of such craft, as 
had been the case with autogiros when helicopters came into being? 

Section 2. The Helicopter Co mp ared to Ve rtical Takeoff 
and Landing and Short Takeoff and Landing 
Aircraft 

When talking of the prospects of helicopter engineering development, one 
must study the problem of the possibility of coexistence of helicopter and 
vertical takeoff aircraft. Do helicopters have a future? Or are the potenti- 
alities of the helicopter exhausted? Can the helicopter successfully conpete 
with vertical takeoff aircraft? Will their development trend teniiinate, as was 
the case with autogiros which ceased to exist with the appearance, in 1940, of 
the first successful helicopters? A conparative investigation of helicopters 
and VTOL or STOL craft as means in transport aviation not requiring an airfield 
will enable us to answer these fundamental problems. 

It is known that recently the matter of vertical takeoff aircraft (in 
English, VTOL) and short-rion aircraft (in English, STOL) has become urgent"'. 
(For footnote, see following page). 

16 



Actually, the present flying speed of fighters, reaching 2500 - 3000 km/hr, 
requires such high-power engines that very little remains to add for their 
vertical takeoff. Therefore, judging by the literature in other countries we 
can assume that fighters and fighter 'bomtiers will be developed mainly as /22 
VTOL aircraft not requiring the use of an airfield. The direction of develop- 
ment of transport aircraft, whose power plant is limited by considerations of 
economics or quite sinply by fuel consumption, tends toward STOL aircraft . 

Some propose that the future development of helicopters will offer a better 
solution to transport problems for a range vp to 600 km than do VTOL aircraft 
or special STOL transport aircraft. 

In examining the possible development trend of aviation, we cannot limit 
the study to an analysis based on the present state of the art in science and 
technology. 

By using such methods, many scientists have repeatedly arrived at erroneous 
conclusions concerning the "limits" in the development of various aircraft or 
helicopters, since they did not provide for the development of parameters' char- 
acterizing the weight and economic perfection of engines or perfection of design 
and materials used. It is necessary to extrapolate their development somehow to 
the future. 

Leaving room in the future for such an investigation, we will estimate the 
situation at hand. We will conpare helicopters with VTOL and STOL aircraft, 
using data of the best helicopters that have been bixLlt as well as of aircraft 
being in the design or construction stage. 

1. Tactica l an d Technica l Require ments for VTOL, and STOL 
Military. Transport^ Ai r craft of the West 

The tactical and technical specifications for VTOL transport aircraft, 
worked out in the USA, call for a flying range of 550 - 700 km, a lift capacity' 
of 3600 kg or 32 troops, and a cruising speed of l+^Q - 550 km/hr at a gross 
weight of not more than 16,000 kg. At the same time a very long delivery range, 
of the order of 4000 km, is required, which is probably intended for the possi- 
bility of ferrying aircraft from the USA over the ocean. 

In studying STOL transport aircraft, one comes across ordinary classical 
propeller transport planes such as, for exanple, the British-Canadian De Havil- 
land "Caribou" (K-g.l.2l). 

By STOL transport aircraft we mean aircraft that use engine power for /23 
reducing the takeoff and landing runs. This is useful and necessary. 

A study of STOL aircraft must include one of the first aircraft of this 
type, the French aircraft Breguet-941 (Fig. 1.22). On this aircraft the entire 
wing area is in the zone of propeller slipstream. All propellers are inter- 



■5!- VTOL - vertical takeoff and landing; STOL - short takeoff and landing. 

17 



connected by a transmission which provides safe takeoff or landing if one or 
two of its four engines fail. The propeller slipstream, deflected downward by 
a double-slotted flap, produces additional lift, which reduces takeoff speed 
and shortens the rtin. However, these qualities are achieved at the e^gjense of 
an increase in enpty weight and shorten the range of this STOL aircraft. Heli- 
copters can operate successfully at such a range. 




Fig. 1.21 British-Canadian Transport Plane 
De Havilland "Caribou". 




Fig. 1.22 French STOL Aircraft Breguet-9/(-l . 



Despite the great type diversity of ?TOL and STOL aircraft, it is not dif- 
ficult to arrange them logically in a general classification of aircraft . They 
should be placed between helicopters and airplanes. 

It is commonly known that the larger the area over which air flows (it 
makes no difference whether it flows through a rotor or the nozzle of a jet 
engine) or, more precisely, the smaller the velocity inparted to the air mass 
for producing lift in aircraft or helicopter, the smaller will be the power re- 
quired for this per unit weight of machine. 

Thus, the ordinary helicopter and the aircraft taking off vertically by the 
thrust of jet engines are at opposite poles of this classification (Fig. 1.23). 



18 



In the pursirLt of greater range and probably higher speed, the helicopter 
was provided with a wing; as the wing area and hence the lift increased further 
(since the thrust of the rotor at maximum flying speed decreases so much that it 
is insufficient for foiTward flight), propellers appeared on the wing. Thus arose 
the British "Rotodyne" (Fig. 1.24) and the Soviet rotocraft designed by N.I.Kamov 
(Fig. 1.25) - aircraft which in place of one lifting and moving system have two, 
one being the rotor and wing for sustention and the other being a system of 
tractor propellers, inclined forward to the thrust vector of the rotor, to pro- 
vide foirward propulsion. During vertical takeoff, the wing and the propellers 
are useless, and in horizontal flight the rotor is siperfluous. The attenpt to 
avoid such superfluous units whose weight unavoidably reduces the useful load 
led to a configuration with a wing and pivoted rotor (Bell XV-3, Fig. 1.26) in /24 
which the rotor in horizontal flight becomes a propeller, and to a configuration 
with a pivoted wing whose propellers during takeoff - turning together with the 
wing - act as rotors as, for exatiple, the XG-142 aircraft produced by Chance 
Vought - I^an - HLller (Fig. 1.27)- 




Jet aircraft taking ofi 
on special engines 




VTOL aircraft 

Propeller aircraft Jet aircraft taking 
with pivoted wing off on fans 
(Miller ^C-i42) ' 





tiotocraft "Rotodyne'* 



Aircraft with short run 
of 130-160 m 



Mi-6 h 




Light transport aircrafti^(Bregiiet-9^1) 



with run of 16ri-:?nn m 
(Car ibou) 



anspor t 
a run 
of 600-800 m (An-10) 




Fig. 1.23 Classification Scheme for VTOL and STOL Aircraft. 




Fig. 1.24 Rotodyne Rotocraft. 



19 



Passing now to aircraft with an engine more powerfiJ. than that of the above 
types of aircraft, the STOL jet aircraft is provided with means for downward de- 
flection of the blast from the jet engines or from various types of auxiliary 
turbofan engines. 

The configuration of the Breguet-941 aircraft (see Fig. 1.22) can be re- 
garded as a variant of an ordinary airplane which, to increase the lift coeffi- 
cient, utilizes the airflow over the wing created by the propellers, or else as 
a variant of an aircraft with a pivoted wing where the thrust of the propellers 
is not literally turned but is deflected downward by means of the mechanized 
wing. 




Fig. 1.25 Rotocraft Designed by N.I.Kamov. 




Z2i 



Fig. 1.26 Bell X7-3 Convertiplane. 



The diameter of the propellers of the WOL aircraft shown in Fig. 1.23 
(from left to right) gradually decreases down to the VTOL jet aircraft which 



20 



II 



has no propeller at all. With a reduction in propeller diameter, the engine 
power increases per unit takeoff weight from 0.25 - 0.3 IpAg for helicopters 
to 3 - 4 1^/kg for jet aircraft (the values of the equivalent horsepower are 
taken here for the aircraft ) . 



Z26 



The cruising speed of these aircraft continuously increases along with the 
increase in installed horsepower. However, this is not a decisive factor for 
the problem of a transport aircraft with a range of 800 - 1000 km. 

This defines the scope of VTOL and STOL transport aircraft to be conpared 
and the flying range over which they are effective. 







Fig. 1.27 Chance Vought - %an - Hiller XC-1^2 VTOL 
Aircraft with Tilt Wing. 



To which of these types of aircraft will belong the future in solving the 
formulated problem? 

Before conparing the helicopter with its competitors with respect to econo- 
my, let us examine the problem of the flying range of the helicopter. In view 
of its conparatively short range, can the helicopter enter this conpetition at 
all? 

Let us first examine and conpare the best of the VTOL and STOL transport 
aircraft that have been or are being constructed: the tilt-wing VTOL air- 
craft of the type XC-142; the STOL aircraft of the type Breguet-941; the regtilar 
transport aircraft of the type "Caribou" HG-4; the rotocraft with turboprop 
engines of the "Rotodyne" type; and helicopters. 



2. Mea ns for Increasing the Flying Range of Helicopters 

The helicopter. has always been regarded as a short-range aircraft; a figure 
of 400 - 500 km is usually given as the maximiim for its normal range. In order 
to treat the helicopter as a conpetitive aircraft in this new area of use, the 

21 



range should be almost doubled while retaining its lift capacity. How does one 
increase the flying range? 



Let us turn to the well-known formula of flying range: 

1 = 270-^^-^ 
G c^ C/ 



121 
(2.1) 



where 



Ce 



G = weight of the aircraft (average during flying time); 
Gt = weight of the fuel; 

= aerodynamic efficiency of the aircraft (taken to be constant); 
= specific fuel consimption of the engine; 

= a coefficient taking into account power losses in the transmission 
due to cooling, etc.; 
Tl = rotor efficiency. 

Equation (2.1) shows that the range is greater, the larger the proportion 
of fuel in the all-up weight of the aircraft and the higher its aerodynamic ef- 
ficiency, engine economy, and effi- 
ciency of engine and auxiliary units . 



This formula holds for any 
heavier-than-air craft, including air- 
planes and helicopters. Specifically, 
it follows from this equation that the 
flying range of various flying 
machines, other conditions being 
equal, does not depend on their cruis- 
ing speed. 



'■jtri 

I 






1 


Airpl 


-r 
ane 


-=j- 


15 
5.0 








>J 


-^'^ 


-^ 


\, 






/ 


^ 










Hel icop ter 






N 




2.5 






^ 






' 















so 100 t50 ZOO 250 300 f-^ 



Can a helicopter be given a range 
sizfficient for coizpeting with STOL 
aircraft? 



Fig.l.2S Product of Aerodynamic 
Efficiency and Rotor Efficiency 
as a Function of Flying Speed. 

As indicated in Fig. 1.28, the 
product of aerodynamic efficiency 
Cy/cx and rotor efficiency 7] for a helicopter with a fixed landing gear is lower 
than for a transport airplane by almost a factor of 2. Furthermore, the fuel 
consunption of the helicopter is somewhat greater than that of the airplane 
since the engine characteristics are inferior at low altitudes and flying speeds. 
Thus, a helicopter can be given a range equal to that of airplanes only by in- 
creasing the fuel svp-plj, i.e., the quantity G^/G. However, in so doing how 
does one maintain the useful load? This can be done only by increasing the 
takeoff weight, but the helicopter will then no longer be able to take off 
vertically. 

What happens if we place these aircraft under equal conditions, i.e., allow 
the helicopter the same takeoff run as an STOL aircraft, namely 150 - 200 m or 
even less? At a relatively large value of Cy, will the helicopter then be able 
to lift - at low speed - a much greater weight than an airplane, accommodate 



22 



more fuel, and thus condensate for its lack in aerodynamic efficiency? 



3. Helicopter with Takeoff Run 



As shown in Fig. 1.29 which gives the curves of the required and available 
horsepower of a "Cari'bou"-type transport airplane and of a modern helicopter, an 
airplane can be kept in the air at a speed not below 115 km/hr. A helicopter /28 

can hover in the air without moving. If 
the heUcqpter is overloaded by 15% above 
the nonnal takeoff weight Gq, it can no 
longer hover and, like the airplane, will 
only be able to fly without dropping if it 
has some speed - in this case, a speed of 
not less than 50 km/hr. At a greater 
speed than this, it will gain altitude and 
at a lower speed, lose altitude. The dif- 
ference here in favor of the helicopter, 
in conparison with the conventional air^ 
plane, lies only in the fact that the 
helicopter retains full controllability 
at a speed below its minimal and that 
there is no danger of separation of flow 
and loss of controllability, both of which 
are possible in the aiiplane. 



»^p 



3000 



zooo 



1000 




Required horsepou 
for airplane 



wo 



200 



300 



Fig. 1.29 Required and Available 
Power as a Ftinction of Flying 
Speed. 



So far as the takeoff distance is 
concerned, assuming that the helicopter 
takes off at a speed of V^in, this dis- 
tance at some average acceleration j, 
will be 



Z.„, 



2 



(2.2) 



Thus, the takeoff vxm is shorter, the lower the minimum flying speed (close 
to takeoff speed) and the greater the acceleration. 



The minimum speed is 



V„ 



V °^- S,nax 



where 



wing area; 
air density. 



What values of c. 



are available to airplanes and helicopters? 



For this, let us calculate the value of Cy that an airplane of the "Caribou" 
type should have at the same weight as the helicopter in order to fly without 
descending at speeds less than minimum. Figure I.30 shows the values of Cy, 

23 



calculated from the formiila 



c.,= 



20 



eSwK2 



(2.3) 



of a helicopter referred to the wing area of an equivalent airplane, which char- 
acterizes the lift capacity of a helicopter in conparison with the airplane. 
The cin-ve Cy of the helicopter in Fig. 1. 30 extends to infinity. This is natural 
since the helicopter has a rotor which in essence is a rotating wing with a 
power plant suspended from it and is capable of producing lift at zero forward 
speed of the entire machine. Here we see that at speeds of 50 - 60 km/hr the 
available values of Cy of the helicopter are several times greater than for an 
airplane of the "Caribou" type at a speed of 115 km/hr, which has a highly 
mechanized wing. 



Thus, at equal power a greater weight can be lifted by the helicopter /29 

at low speeds than by an airplane. 

However, a greater flying weight 
does not always mean a greater useful 
load. 

At equal relative fuel weight 
(about 10^), the ordinaiy airplane of 
the "Caribou" type has a range of 
1000 km, i.e., twice that of a heli- 
ccpter taking off without a run. 

The Breguet-941 STOL aircraft (at 
a fuel weight 12 - 13^ o'f the flying 
weight) has twice the range of the 
helicopter or of the X6-142 VTOL air- 
craft . 

If, in helicopters, the fuel weight 
is increased to 20 ,- 25^ of the gross 
weight, then the range of the helicopter 
can be doubled and raised to 1000 km. 
This value is already close to the 
normal ranges of specially designed 
STOL aircraft. 



(■''he 


1 












10.0 






























8.0 

























BJO 














\ 








U.0 
(<■ 


\ 






\ 


^ 




2.0 




\ 






1 


\ 




^l^'lLin 


,1 . 
aire 


Biin 




^ 





(G=WGg] 100 



200 



hr 



Fig. 1.30 



Dependence of (cy ^j), 
on Flying Speed. 



sq 



The load ratio of helicopters 
taking off with a imn and at increased 
fuel sT^jply becomes higher than the load 
ratio of coiiparable aircraft and reaches 
44 - 50^. This makes it possible 'to ob- 
tain equal productivity at almost the same takeoff weight of airplane and heli- 
copter. For exanple, a transport helicopter of average lift capacity, just as 
a "Caribou"-type airplane, can transport a load of 3.2 tons over a range of 
1000 km. It is true that the helicopter, in so doing, uses 2.5 times more fuel. 
However, it must be remembered that the airplane needs twice the area for taking 



24 



off and, what is quite inportant, the helicopter after having constimed half its 
fuel is able to land vertically, whereas the ordinary aiiplane cannot do so. 

It must be enphasized that conparable aiiplanes and helicopters have 
practically the same power supply (0.23 - 0.25 Ip/kg). One must also bear in 
mind that piston engines, operating on gasoline, have a lower fuel consunption 
at low altitudes than turboprops, so that the average tvirboprop helicopter oper- 
ates under less advantageous conditions than the "Caribou" aircraft with piston 
engines . 

Thus, the suggestion to use a takeoff run for the helicopter will permit 
doubling its range at the same useful load. 



4« Takeo ff Distance of Helicopter 

We have already expressed the takeoff distance in terms of takeoff speed 
and acceleration. The takeoff speed, proportional to the minimum speed at which 
a helicopter can be supported in the air at an overload of 15^ as opposed to /30 

the weight with which it can take off 
without a run, is not more than 
"" '"'"-- - 60-70 km/hr. Let us now define the 

possible degree of linear accelera- 
tion, since the takeoff run is in- 
versely proportional to acceleration. 
Let us find the possible initial ac- 
celeration. 




T«-T*T 



nose wheel 



tail wheel 



As agreed, let the helicopter 
develop a thrust amounting to only 
0.85 (takeoff weight) at the take- 
off power. Then, allowing for some 
angle of inclination of the rotor 
axis to the vertical Sf (here the 
difference in the compression of the 
struts and pneumatic tires of the 
nose and tail wheels is accounted for) and for the forward deviation of the re- 
sultant owing to deflection of the automatic pitch control mechanism through an 
angle Dik, according to Fig. 1. 31, we find the initial acceleration: 



Fig. 1.31 Forces Acting on Helicopter 
during Takeoff Run. 



^0= 5-[-^sin(D,x-;-£^)-^l--^yj, 



(2.4) 



Here, the second term on the right-hand side takes into accoiont friction 
of the wheels against the ground, with a friction coefficient f . Adopting the 

T 
usual notations of e, = 6.5°, DiH = 10°, -^ = 0.85, and f = 0.12, we obtain 

Jo = 2.2 m/sec^. 

p 
Assuming a relative static propeller thrust of -rj- = 1.6 kg/tp for the 



25 



airplane and us5 .g = 0.25 i^/'i^gf the initial acceleration -will be 

G 



/n=fi'( /W2.5 m/sec^ 



Jo^g 



\N G 



i.e., a value of the same order as for the helicopter. 

Of course, the acceleration at the moment of takeoff is determined by the 
excess power, which is somewhat higher for the airplane. However, its propeller 
thrust decreases with an increase in speed whereas the rotor thrust increases; 

in fact, the angle of pitch of the 
heUcqpter, during the takeoff 
run, may even increase since, 
during takeoff, the tail wheels 
are able to Hft off the ground 
at a thrust substantially less 
than the takeoff weight so that 
the takeoff run is conpleted- on 
the nose wheel. 




Fig. 1.32 Forces Acting on a Helicopter 
during Takeoff on Nose Wheel. 



(-0) 



-1 

takeoff 



It is obvious from Fig. 1.32 
that the thrust-to-weight ratio at 
which the tail wheels can lift off 
the ground (disregarding friction) 
will be 



(2.5) 



where x„ax is the distance from the center of gravity to the axis of the auto- 
matic pitch control mechanism at a maximal forward deflection/ of this mechanism 



[here it is assumed that the quantity 



x„ 



can be neglected for unity] . 




Fig. 1.33 Running Takeoff of Mi-6 Helicopter. 



26 



With the usual relations, this corresponds to a thrust-to-weight ratio of 
0.8 - 0.85 • Figure 1.33 shows a helicopter during the takeoff run, at a gross 
weight of G„ax = 1.15 &• 

An exact calculation of the takeoff run can be carried out by the same /31 
method proposed by the author 30 years ago for calculating the takeoff run of an 
autogiro ( Ref .4) • 

Running takeoffs perfonned in practice have confirmed that, at a 15^ over- 
load of a helicopter as opposed to the maximal weight with which it can take off 
without a run, the takeoff run amounts to no more than 60 - 100 m in still air. 

5. Criterion for Estimating the Economy of Various 
Transport Aircraft 

In any conparison of two transport aircraft, attention is primarily centered 
on the lift capacity. Still, the speed of transport is also important . Actual- 
ly, if a load can be transported more quickly, then more loads can be trans- 
ported in unit time over a given distance at a smaller lift capacity. 

This results in the well-known criterion of hourly productivity Gio^d^av "t ' 
• km./hr (Y^^ is the average ground speed). 

However, at what cost is the load transported? 

If both aircraft have identical efficiency and range as well as takeoff /32 
and landing properties satisfactory for fulfilling the mission, which should be 
given preference? 

To answer this question we must know which of the aircraft is more economi- 
cal. In military use, the advantages of any aircraft for solving transport 
problems, which sometimes arise at an appreciable distance from the supply 
bases, are determined primarily by cost data. Expenditures for construction of 
the machine itself, incurred in the past, are no longer of significance and have 
no effect on fulfilling the immediate task. Under such conditions, the economy 
of an aircraft is determined mainly by the amount of fuel consumed. Here, the 
transport of fuel constitutes a bottleneck that is decisive for the ability to 
solve the stated problems . The criterion of economy under such conditions is 
conveniently obtained by referring the hourly productivity to the weight of the 
fuel consumed during that time Gtiji: 

^..= ,^. (2.6) 

Since the fuel cons^l^ption per kilometer is 

^ (2.7) 

it follows that 

27 



The quantity L^^g has the dimension of length, so that we can call it the 
equivalent specific range of the aircraft. It represents the distance over 
which a given aircraft can fly in excess of the design range if the entire trans- 
ported load is replaced by fuel. Still another meaning can be given to this 
quantity. It can be regarded as the distance over which an aircraft can carry 
one ton of cargo after having consumed one ton of fuel. It is clear that the 
quantity L^.g depends on the distance of transportation just as productivity 
depends on it. The farther the machine flies, the more fuel it needs and the 
smaller the cargo it can take at a given flying weight (maximal). 

On the other hand, L^^g is the work expressed in ton-miles which a given 
aircraft can perform, having consumed one ton of fuel. 

The inverse quantity of L^ ^ , i.e., — = , is fuel consunption in tons 

required for performing transportation work of 1 ton-mile or kg/ton-mile. 

We can also use other criteria that estimate economy, i.e., the cost of 
transporting one ton-mile in rubles. In these criteria, we can take into ac- 
co\ant the cost of the aircraft (to some extent, this is proportional to the 
entity weight of the aircraft), the service life of the conponents and power 
system, cost of operation and repair, etc. 

In this problem, the military transport criterion L^^ ^ or — — - — = Cl is 

most iirportant since it takes into account not only econony but also the real 
and ponderable requirement of sijpplying fuel for transportation equipment under 
military conditions. Therefore, we can disregard roany other criteria but not 
this one. 

We sipplemented the values of L^^g and Cl for conparable aircraft by /33 
data on jet VTOL aircraft which, for hovering, use specially installed lifting 
jet engines or which take off by means of fans driven by the main engines. The 
fuel consunption of such aircraft, while hovering, constitutes such a large per- 
centage of the takeoff weight that it must be taken into account when calculating 
the value of Cl ; therefore, the corresponding formula takes the form 

where iqOO G, 

Cl • = -p j^— is the fuel consunption xn horizontal flight; 



"h O V 



Cl = 1000 — ■- is the fuel consumption while hovering. 



G 

For an aircraft taking off by means of wing fans rotated by turbines 
mounted to their periphery, we have used data similar to those published for the 
experimental Ifyan X-16 airplane constructed in the USA. 

In the calculations, we disregarded the fact that the installation of the 

28 



lifting engines inpairs the aerodynamic efficienciy of a VTOL aircraft. It was 
assumed that, at ranges greater than 1000 km, VTOL aircraft of the indicated two 
types with a sustainer turbojet will fly at the design altitude at maximum aero- 
dynamic efficiency, whereas at a reduction in flying range from 1000 to 50 km 
the operational ceiling decreases accordingly and any drop in aerodynamic effi- 
ciency leads to some increase in fuel consuoption. The hovering time t^ov was 
taken as 6 min (3 min in takeoff and 3 niin in landing) . 



Table 1.3 gives owe calculated data for aircraft of different flying ranges 
and different forms of takeoff. 











TABLE 


1.3 


















Vertical 


Typ e o f Tak ec 

With Short 
Run 
i, <200 m 


ff 

From Small 
Airfield 
I, <400 m 




Fr"; 
Con 
Air 






Characteri sties 


)m 

:rete 
strip 
«8D0m 


Road 
Transporta 


Range t,. km 




■ 500 


1000 


1000 


2000 


_ 


Flying altitude 


V 

a. 

o 
u 

u 
X! 


500-2000 


500-2U0O 


3000 


8000 


- 


Typ e o f 
fli rcraft 


VTOL 

Propell er 

' Aircraft 


4-> 

(0 
U 

o 


o. 
o 
o 

■ H 


STOL 

Propeller 
Aircraft 


CO 

t-< 
o 

2 

£ 


Light Transport 
Propeller 
1 Aircraft 


1 .STOL 

Propeller 

Aircraft 
(Overload Weieht: 


[ Turboprop 
Transport 
Aircraft 


Turbojet 

Transport 

Aircraft 


° s 

[- a: 


a 
a 

I 3 

m u 
«i- 

o 
u 

o 


Takeoff run 
I , m 











80 


140 


1£0 


170 


210 


850 


950 


— 


— 


't' 
C^, kg/tkm 


0.8 


0.9 


1.3 


0.9 


0.6 


1.4 


0.4 


0.5 


0.3 


0.5 


0.13 


0.4 


V^^, km/hr 


250 


460 


280 


240 


400 


280 


250 


400 


600 


800 


60 


30 



To conpare the economy of transport conveyances, the values of Cu are given 
for the Soviet truck ZIL-I5I where it is assumed that the road between two /34 
points is longer by a factor of I.5 than the air route. 

Figure 1.34 shows the values of Cl as a function of flying range, for vari- 
ous VTOL aircraft. The value of the flying range L in calculating Cl was deter- 
mined with consideration of a decrease in fuel consunption per kilometer as the 
aircraft became lighter due to depletion of the fuel. Assuming a linear de- 
pendence between consunption per kilometer and the weight of the aircraft, we 
must correct the value for the range, calculated from the above formulas where 
the per-kilometer consunption is accepted to be constant and corresponding to 
the takeoff weight, by the quantity 



Ki= 



log 






29 



I t 



where _ q 

G-r = — 7=r— is the ratio of fuel weight to takeoff weight of the aircraft. 

For TTOL aircraft, aircraft with a large fuel consunption for hovering, the 
reduction in flying weight due to the expenditure of fuel for hovering and in 
horizontal flight was taken into account. Figure 1.35 shows the curve of the 
correction coefficients Kl as a function of the values of Gj. 



C^ kg/ ton- km 




VTOL turbojet 
+—— a.ircr aft 

Lk 



Fig. 1.34 Dependence of Cl on Flying Range L. 



For aircraft with takeoff rim, the values of Gl at different fljring ranges 
are given in Fig. 1. 36. The diagram also shows how the economy of transport 

means can be increased, at a given 
range, by using a takeoff rion. The 
f(^ longer the takeoff, the larger the take- 

off weight of the aircraft and hence the 
greater the weight of transported cargo. 

Such are the results of investigat- 
ing fuel consunption for transporting 
one ton-mile with various types of 
transportation means. 

As regards the cost of operating 
airplanes and helicopters, which natural- 
ly is determined not only by the cost of 
fuel but also by the service life and 
initial cost of the machine, we must bear 
in mind that the greater power/weight /36 
ratio of the VTOL aircraft conpared to 
that of helicopters as well as the 
presence of transmissions in some types 
more or less balances this cost. As for safety in the case of engine failure 



a 

V 

in 








y 






y 


— - 









0.1 



0.1 



0.i 



O-hG^ 



Fig. 1.35 Dependence_of Coefficient 
Kl on Gf . 



30 



diiring takeoff or landing, all advantages are here on the side of helicopters 
since the propellers on an aircraft -with a pivoted vjing are not capable of auto- 
rotation, and engine failiire (depletion of fuel) during landing of an aircraft 
of the type Breguet-941 may lead to flow separation and an uncontrollable de- 
scent because the wing is no longer washed by the propellers. Furthermore, 
there will always be the difficulty of providing controllability at low speeds 
in these machines. 

C, kg/ ton-km 

~f^VTOL turbojet 



15 



as 



I Q-ircraft 




Cros s-Coun try 
Truck 

wo 



/YTVTOL propeller 

aircraft STOL propeller 

^^--J____^SS^-^ aircraft 

- I — ^' ' 

We I icop ter 



Range 
2000- 3000 kxf, 



I. ight transpor t 
P 



.Turbc 



aircraft ^^fe^ _J ":''"^''°/' 

. -.■ aircraft 
t rack on highway 



ZOO 



m 



m 



500 



Lr m 



Fig. 1.36 Dependence of C^ on Takeoff Run for 
Various Flying Ranges. 



A further reduction in weight and fuel consunption of turboprop engines 
under development at present will lead to an increase in load factor of heli- 
copters. The substantial increase in service life of rotor blades, reduction 
gears, and transmissions obtained in modern prototypes will equalize the amorti- 
zation cost of airplanes and helicopters, after which the helicopter will become 
a full and equal member of the air transportation system in its most massive 
area. 

6. Possibilities of Increase in Maximum Flying Speed 

If the flying speed of VTOL and STOL aircraft is considered to be an int- 
portant flying and tactical requirement, then the possibilities of rotocraft are 
far from exhausted with respect to further increase in speed. 

If we equate the power required for horizontal flight and the net power of 
the engine, we can obtain the relation between maximum speed and power/weight 
ratio of the aircraft: 



^'„,.. = 270 



N 



11^, 



(2.10) 



31 



where 

M = engine power; 

G = gross weight of the aircraft. 

It follows from eq.(2.10) that the maxiinian speed is directly proportional 
to the power/weight ratio N/G of the aircraft. 

Pigure 1.37 gives the curves of the required power/weight ratio as a func- 
tion of flying speed for various aircraft. The curves for heavy rotocraft show 
that, to increase the flying speed above 300 - 320 km/hr, it is necessary to 
sipplement the helicopter rotor with a second high-lift device - a wing; to 
reach speeds above 370 km/hr also propellers are needed, which means changing 
over to a rotocraft. Thus, by formulating the problem of achieving the highest 
speed possible at any price, it becomes possible to decide what VTOL aircraft 
config-uration to use for different majdmum flying speeds. However, it must be 
remembered that the transition from heUcopter to rotocraft involves a loss in 
1 T ft capacity, an increase in the cost of construction, etc. Even with a 
power/weight ratio of 0.45 J^'/kgj which can presently be realized on rotocraft, 
the transition from helicopter to rotocraft will not produce a gain in speed by 
more than 30 - 40 km/hr. 




/37 







WO 200 M 1*00 500 600 W V km/hr 



Fig. 1.37 Power/Weight Ratio of Aircraft as a 
Function of Flying Speed. 



Finally, the graph of the required power/weight ratio clearly shows the 
great difference in the power/weight ratio of VTOL aircraft and of rotocraft. 
At equal power/weight ratio, the rotocraft is somewhat inferior in speed to the 
prcpeller-driven VTOL airplane with a short range. 

32 



Section 3- Basic Principles of Design 

1. Selection of Engine Horsepower and Rotor Span 

In most cases, the helicopter designer is siJpplied with the desired lift 
capacity. Knowing the required speed, he estimates the necessary power/weight 
ratio. After assigning the current percentage of the useful load ratio, he de- 
termines the order of magnitude of the flying weight and hence the magnitude of 
the installed power. Having selected the number of engines in view of the end 
use of the helicopter (one engine for a light military machine, at least two 
engines for a passenger craft, etc.), he can select the most siu-table engines 
among existing or scheduled types. 

Usually, it will happen that the power of the possible combinations of 
engines does not match the desired power. This necessitates correcting the para- 
meters of the helicopter in question, after selecting the optimton combination of 
existing engines. After this, the main problem facing the designer is to select 
the rotor span for the specific power plant. 

How does one select disk loading? 

It is known from statistics that disk loading rapidly increases with inr- 
creasing flying weight and varies within 12 - 50 kg/m^ as the weight increases 
from the lightest to the heaviest helicopter. 

Disk loading, as a function of weight, varies even more than wing load- /38 
ing of an airplane. This is of inportance since an increase in wing loading of 
an airplane can be conpensated by an increase in length of takeoff run whereas, 
for helicopters, the takeoff run must always remain zero. 

The weight of the rotor increases approximately in proportion to the cube 
of its span. However, at equal power the lift capacity of the helicopter as a 
whole increases in proportion to the 2/3 power with an increase in span. In 
addition, such flight data as ceiling, rate of climb, range, rate of descent in 
autorotation must be inproved when the span is increased and hence disk loading 
is decreased. 

It is inpossible to calculate the parameters of an optimal design since 
there are too many contradictory considerations that the designer must weigh. 
The answer to this problem should also include a search into the past, an analy- 
sis of the development of helicopters with respect to size. 

How is the next (larger) helicopter to be developed? 

It is obvious that the prime requisite is to increase the installed power. 
However, to what extent? For exanple, if we retain the power/weight ratio (for 
considerations of economy) and then, if necessary, increase the rotor span so 
that the former disk loading remains, or else if we increase the power/weight 
ratio and then have the opportunity to increase disk loading provided that take- 
off is vertical, will we actually obtain a conparatively smaller rotor span? 

The best solution is obtained with the variant having the lowest construction 

33 



weight, i.e., providing a higher proportion of usefiol load. 

Let us find the variation in load ratio of a helicopter when its size is 
increased for different power loading and disk loading. 

We will examine the case in which a new larger helicopter is developed ac- 
cording to the same scheme as above. 

Taking the static thrust (without consideration of the ground effect) to be 
equal to the weight, Joukowski's formula will furnish 

r=G = (33.25;iiA^D)2/3, (3.1) 

where 

g = a coefficient smaller than unity characterizing the mechanical losses 
of power in the transmission and those due to cooling and equaliza- 
tion of the torque; 

1] = rotor efficiency representing the ratio of the useful power needed 
for s'upporting the aircraft in the air during hovering to the spent 
power. 

We will write the e:!^ression for takeoff weight of a helicopter in the form 

G=Ouse +G„L + A,,7V + 0,,G, (3.2) 

where 

Gygg = usef-ul load of the helicopter; 
G„i = weight of the nonlLfting elements of the helicopter; 

G = takeoff weight of the helicopter; 
ktr = an enfiirical coefficient representing the ratio of engine weight 
_ and transmission to engine power; 

Gjs = relative weight of lifting system (rotor with hub and automatic 
pitch control mechanism) with respect to helicopter weight. 

Change in load factor on increase in rotor span . We will attenpt to de- /39 

n. 

termine the load factor — "!," ■ of a heavier helicopter which differs from the 

G 

original by the rotor span: 

Assuming that, in this case, the power/weight ratio N/G remains constant, 
eq.(3.l), at a constant value of §T1, will yield 

G3/2 = const A^D 
or 

34 



"1/ -z^ = const. 



G_ f G 

nVd^ (3.4) 



Hence the disk loading, i.e., the quantity 

G 
P= ■ 

JlD2 

4 
also should remain constant. 

On increasing the rotor span, if the size of the blades changes similarly 
while their number remains unchanged, the weight of the lifting system increases 
in proportion to the cube of the ratio of the rotor spans: 

G,,, = Gu.(|-;. (3.5) 

Consequently, 

Here the subscripts 1 and 2 pertain, respectively, to the original heli- 
copter and to the helicopter under study (Dg > Di). 

However, if the disk loading p is increased, flow separation at maximum 
speed can be avoided only by increasing the loading (mainly by increasing the 
n-umber of blades since a relative increase in chord is less advantageous and 
causes a greater increase in weight because of the need for larger balancers to 
eliminate flutter) . 

In this case, when retaining the span and the tip speed of the blades, the 
weight of the lifting system will increase proportionally to p, i.e., 

' PiKoJ (3.7) 

or 

Gi.s, = Oi,s.|^. (3.8) 

Substituting this expression into eq.(3.3)» we obtain 

G.S., =l-5nC-V(-f),-^^-lr- (3.9) 

The coefficient k^r is the sum of the relative weights of engine and trans- 
mission: 

*tr=-^ + ^^- (3.10) 

35 



mill I II I 



The first addend remains unchanged (it characterizes the weight charac- /k.0 
teristic of a modern engine) and the second increases proportionally to the in- 
crease in rotor span. 

Actually, if we assimie that the weight of the transmission is proportional 
to the magnitude of its transmitted torque M^., then we obtain 

^^--^'^'m; '''''^- (3.11) 

Keeping the tip speed constant (uoR = const), we have 

•'I ^ D2 
"2 Di 

Hence, 



At constant disk loading it follows from eq.(3'4) that 



where 



Then, considering also that 



and 



ffi =02/^1- 



<J2 Gi 



(3.12) 



(3.13) 



we can transform eq.(3«9) into 

^"" =>-°.^-T (^-[°- +t(^)J^' (3.15) 

Where ^^ ^ ,0> 

— P G 

Assuming for the original version: Gni = 0.25; — ^^^- = 0.2; — ^ = 0.4; 

Gi = = 0.18: -^ = 0.28 and substitiifcing these into eq.(3.15), we obtain (at 
G2 = Gi) 

36 



^G^ts. — \ =0.402; 



G /I 
Ous, =0.694-0.292,)/^. 



(3.16) 



Hence it is clear that the magnitude of the load ratio of a larger heli- 
copter decreases monotonicaUy vdth an increase in m (Fig.l.38). 

The ratio of the useful load of helicopters can be represented as 

G, 



f Guse \ 



o 



( G use \ 
a A 



= 2.49/ra{0.75 L^.^ 



-["•'«+^m]^^)- 



(3.17) 



It follows from eqs.(3-15) and (3-17) that it is inpossible to construct /hi 
a larger helicopter while maintaining the same disk loading and the same power 
to weight ratio as those of the original helicopter, with a larger (or even the 
same as the previous) coefficient of the load ratio, although the absolute value 
of the useftil load increases at first (Fig.l.40). 















p= const 








Ouse Original 


hel icopter 


0.3 




~~- 




-^ 


^ 













0.1 














1 


~rr 




-^ 


0.1 
















- 


— 






n 























OM 



D' const 

Original helicopter 




Fig. 1.38 Dependence of Load Ratio on Fig. 1.39 Dependence of Load Ratio on 
Scale m of Weight Increase when Scale m of Weight Increase when 
p = const. D = const. 

Charge in load ratio on i ncrease in power/weight ratio without change in 
rotor span . Let us next examine the case where the rotor span remains unchanged 
while flying weight and engine power increase, i.e., the power/weight ratio of 
the helicopter and disk loading increase. 

It follows from eq.(3'l) that the power/weight ratio of a helicopter should 
in this case increase according to the law 



aXHi).^ 



(3.18) 



37 



Equation (3.8) indicated that the relative weight of the lifting system re- 
mains constant. Keeping the ratio of engine weight to engine power constant, 
the relative x^eight of the engine increases in proportion to /m 



"'-H-^Xi >- 



m. 



(3.19) 



The relative weight of the transmission also changes 



O. 



tr 



V ^ A 9. 



Ym. 



(3.20) 



Then the expression for the load ratio coefficient will take the form 



G 



■A 



'or 



Gus- 



£tng 



N 



N 



9l 



(3.21) 



Substituting the numerical values, we have 

G use =0.57-0.17 Vm- 



The ratio of useful load to helicopter weight varies in the same manner as 
in the previous case (see Fig. 1.39), i.e., decreases monotonically. 

As we see (Fig.l.40), upon an increase in power at constant span, the use- 
ful load increases more rapidly than upon an 
increase in span at constant power/weight ratio. 

Thus, it is obvious that the disk loading 
depends on the weight characteristics of /42 
specific engines available to the designer for 
solving the formulated problem, namely to lift 
a prescribed useful load. It is obvious that 
the lighter the engine in conparison with a 
given prototype, the greater will be the optimum 
disk loading and the smaller will be the rotor 
span. 

This is the reason for the small loads per 
square meter of the first airplanes and heli- 
copters. These machines ■with their then low- 
power and high-weight engines with large load- 
ing were generally not airworthy. 




Fig. 1.40 Ratio of Useful 
Load Weights as a Function 
of Flying Weight Ratio. 



Thus, the designer or researcher who 
■wishes to project into the future should adopt 
some rules for decreasing the unit weight of 
engines, rotors, and nonlifting structural 
elements (by using new materials and increasing the effective design stresses) 



38 



aside from the possible discovery of new engine operating principles; only then 
will he be able to predict the potentialities of developing larger or more 
economic (load ratio, flying range) aircraft. 

It woiild be natioral to e^q^ect a pronounced increase in load ratio by re- 
turning to small helicopters and using the level of engine and rotor unit 
weights achieved in developing the heavy helicopters of the 1960s. 

Actually, a conparison of recent helicopters with turboprops of the same 
weight category as the Mi-1 and Mi-4 helicopters showed that their load ratio 
almost doubled. 

2. Analysis of Multirotor Configurations 

Sooner or later, the designer is confronted with the problem of the expedi- 
ency of further increasing the rotor span and the need to change to a twin- or 
multirotor configuration. 

The lot production of still another blade size requires very large capital 
investment for b\jilding new steel mills, presses, and other expensive equipment 
required for finish-treating of spars and blade assembly. Therefore, the de- 
velopment of new blades is to some extent a Federal problem. At the same time, 
termination of the production of any one type of series-produced blade is im- 
possible, since the existing inventory of helicopters is a steady consumer of 
blades because the blade life, as a rule, is considerably shorter than the 
service life of helicopters. Consequently, when initiating a new blade design 
new production facilities must be created to supplement those already available. 

Therefore, after having developed the largest series-produced rotor, it is 
logical to attack the problem of the optimum multiple to be used. This renders 
the problem of configuration specific: it becomes necessary to double or triple 
also nonrotor units, i.e., rotors together with reduction gears and engines. /43 

Actually, the number of combinations is not excessive: twin- rotor (side-by- 
side and fore-and-aft configurations) and three- rotor helicopters. The cimiber- 
some four-rotor configuration need not be discussed here since the above con- 
figurations are able to provide the maximum required lift capacity of 40-50 tons. 
Another problem to be discussed is that of conparing single- rotor helicopters, 
designed for similar missions, with these configurations. 

Fore-and-af t_ c£nf igioration . Since the induced velocities of the front and 
rear rotors are identical, the induced velocity of the system will differ only 
by the quantity of the average velocity of mutual induction 

'"<^^=T ■'■'"" (3.22) 

where 

vi = induced velocity in the rotor plane; 
K = coefficient of induction. 

Then, the additional induced power of the system or of the rear rotor is 

39 



aA^,=-— y.v.G. 

i 2 • 



(3.23) 



If both rotors lie in the same horizontal plane and do not overlap, then 



= 2; at -^ = 0.2 (Fig. 1.41) we already have h = 1.35. 



However, we shoiild consider that the induced velocities are unevenly dis- 
tributed over the disk so that the average ve- 
locities are larger, corresponding to another 
aspect ratio, i.e., to a smaller span or, what 
comes to the same, to a larger load p on the 
sipporting surface. Therefore, we will take the 
expression for the average velocity of mutual in- 
duction, referred to the entire craft, as 




v^^=0.75vi=0J5 



4p 
VA 



(3.24) 



Fig. 1.41 For Determining 
Rotor Overlap . 



Then the rate of climb of a helicopter of a 
fore-and-aft configuration can be e^qjressed by 
the rate of climb of a single-rotor helicopter 



y 



-V^ 



3p 



(3.25) 



where 



Vy 

Vy = 

A = 



rate of climb of a tandem (fore-and-aft configxjration) helicopter; 
rate of climb of a single-rotor helicopter; 
relative air density. 



Thus, whereas the Mi-4 helicopter with blades of mixed design has a vertical 
rate of climb, at a flying speed of 100 km/hr, of 3-6 m/sec, the vertical speed 
of the tandem-rotor helicopter (Fig. 1.42) with two such power plants decreases 
by the quantity 



3-21 



AVt=l/^-l/t ^'^:£i=,2.25 m/scc, 



28 



i.e., the rate of climb of the tandem^ rotor helicopter is 1.35 m/sec. 

Consequently, the flying characteristics of the fore-and-aft machine sub- 
stantially differ from those of the original single-rotor helicopter from which 
the power plants were taken. 

The rate of climb is determined by the transverse span of the helicopter, /l^U 
its engine power, and the takeoff weight of the helicopter prescribed by the 
designer. It can be stated that, if these parameters are given, the maximum 
possible rate of climb will be determined regardless of the configuration of the 
planned helicopter and the type of its power plant. 

In addition to power e^qp ended for lifting weight, there are also power 



40 



expenditures for mechanical, profile, and induced drag. Thus, if the takeoff 
weight of the helicopter of a fore-and-aft configuration is twice the weight of 
a single-rotor helicopter, the magnitude of rate of climb can be maintained only 
by increasing the power of the tandem-rotor helicopter to more than the double 
power, by an amount of 

^ 75K4 

If we only double the power of the tandem-rotor helicopter, its rate of 
climb will decrease, as indicated above. Such a doubling would be especially 
unsuitable for heavy helicopters with large disk loading; thus, at values of 
p = 40 kg/m^ and V = 40 m/sec the loss of rate of climb AVy at an altitude of 
H = 3000 m, in conparison with the original single-rotor helicopter, will be 
4 m/sec, i.e., such a helicopter will not be able to fly if Vy < 4«5 - 5.0 m/sec. 




Fig. 1.42 Yak-24 Helicopter. 

Cpnparative data of two Soviet helicopters Mi-4 and Yak-24 of different 
configiorations but having the same engines and supporting systems (these are 
doubled on the Yak-24 helicopter) show that the fore-and-aft helicopter has a 
vertical groiind speed 2.6 m/sec lower than that of the single-rotor helicopter 
and that, at altitudes of 1000 - 2000 m, the loss in vertical speed reaches 
3-3*5 m/sec. The service ceiling also drops by a factor of 2 for the fore- 
and-aft helicopter. 

Figiare 1.43 shows the change in torque distribution with respect to the 
rotor shafts, measured in flight on one of the fore-and-aft helicopters. At 
p, = 0.1 - 0.25, the rear rotor consumes about double the power of the front 
rotor. 

This nonuniformity in loading of the rear and front reduction gears and 
rotors substantially reduces the lifetime of the rear rotor parts or else neces- 
sitates development of a more powerful and heavy reduction gear. It is inpds- 
sible to use the main reduction gear of the original single-rotor helicopter as 
the rear reduction gear of the tandem helicopter. 

The large induced losses due to mutual interference of the rotors in fore- 
and-aft helicopters, which amount to 20 - 25^ of the power at cruising speed, /45 

41 



greatly impair its efficiency in conparison with the single-rotor helicopter in 
which the power expenditiire for driving the rear rotor in horizontal flight is 
negligible . 



Side-by-side configuration . This 
old problem of how to build a "bridge" 



/"t 


_^ 












0.0015 




\ 


Two ro 


tors 






s 


V 


• 


• 




0.0010 












_-- 


>< 


Re 


jr roto 


-■* — 

r 


• 


0.0005 




\ 






'i*""^ 






Fron 


>- 

t rotoT 





















configuration is another solution to the 
connecting the rotors . How can one de- 
fine the difference between such fore- 
and-aft and side-by-side "bridges"? 

From the viewpoint of mutual inter- 
ference during hovering, there are no 
fundamental differences in the operation 
of the rotors of either type of heli- 
copter configuration. However, the 
losses from propeller wash over the wing 
in a side-by-side helicopter may be 
greater than from wash over the fuselage 
in a fore-and-aft helicopter. There- 
fore, the characteristics of vertical 
takeoff are poorer for the side-by- side 
configuration with a wing. Neverthe- 
less, in forward flight the rate of 
climb of the side-by-side helicopter in- 
creases by an amount of 



Al^y = — y-iJi^ 0.3751), = 



1.5p 
VA 



0.1 



02 



Fig. 1.43 Change in Torque Dis- 
tribution with Respect to Rotor 
Shafts as a Function of their 
Operation. 



in conparison with the original single- 
rotor helicopter, due to a decrease of 
induced drag (the "span" of the side-by- 
side helicopter being twice that of the 
original single-rotor machine). 



It can be demonstrated that, between 
the load ratio of a side-by-side helicopter and a single-rotor helicopter, the 
following relation exists: 



where 



5g = 







V G 


5r e ar 


= 


Ne„. 






5t 


= 


Tl 


_ Gw 


- 


G-rear 



(i-Lr[(-^^)^^-^.-m. 



Tl 



a coefficient taking into account power expenditures 

at the rear rotor in a single-rotor helicopter taken 

equal to 0.09; 

a coefficient taking into account thrust losses due to 

wash over the wing, taken equal to 0.07; 

a coefficient taking into account the variation in 

weight on changing from a single-rotor helicopter to a 

two-r*otor helicopter of side-fcy-side configuration. 



Let us assume that the wing weight is 12^ of the weight of the single-rotor 



42 



helicopter and that of the rear transmission mth rotor, 10%. 

The above formula shows that, in a side-by-side helicopter constructed from 
two single-rotor helicopters, it is inpossible to achieve the load ratio of the 
single- rotor helicopter. 

If a side-by-side helicopter is designed from scratch, it might result /U6 
in a much better construction since the designer of a side-by-side helicopter 
will not use the same load per square meter as the designer of a single- rotor 
helicopter had been using. In fact, this factor will be made greater and the 
helicopter will become more conpact. On the other hand, when designing a fore- 
and-aft helicopter from scratch it is natural to select a smaller load per area 
and a higher power/weight ratio than is usually done. 

It must be borne in mind that, in designing a side-by-side helicopter, 
there must be means available for "controlling" the frequency of natural oscil- 
lations of the wing with power plants and rotors, both in the vertical and hori- 
zontal plane, since this configuration has a multitude of possible vibration 
modes whose frequencies may enter into resonance with the forced frequencies in- 
duced by the rotor. Furthermore, in the case of high-power and thus heavy 
engines moimted to the wing tips, the side-by-side helicopter almost certainly 
will have a vibration mode of a frequency close to or even smaller than the 
rotor rpm, at which the rotors will vibrate on horizontal displacement. This 
may set tp oscillations of the "ground resonance" type not only on the ground 
but also in the air. Therefore, the designer who has decided to design a side^ 
by-side helicopter is faced with the difficult task of making the wing as small 
as possible in area, light In weight, and sufficiently rigid in bending and 
torsion. 

However, despite certain difficulties in designing side-by-side helicopters 
and shortcomings of fore-and-aft configurations, designers will have to resort 
to them as a means of increasing lift capacity. This becomes obvious when con- 
sidering the difficiilties In developing rotors and reduction gears of super- 
heavy single- rotor helicopters . 

Selection of configuration . An analysis shows that, in changing over to 
side-by-side or fore-and-aft helicopters with doubling of the power plants, it 
is lirposslble to double the useful load lifted by the single- rotor helicopter. 
If this could be achieved at all, it woiild be at the price of an appreciable loss 
in such flying and tactical data as takeoff and landing properties, rate of 
climb, dynamic celling, etc. Thus, the transition from the single- rotor to the 
multirotor helicopter must be done over an increase in power/weight ratio. 

However, the selection of the conflgiiration can be largely influenced by 
factors such as end use of the helicopter and tactical, technical, or operational 
requirements. The designer often prefers to adhere to the configuration for 
which he has more data and ejjqaerience, if other conditions permit selecting 
several approximately equivalent configurations. In some cases, the designer is 
forced to give preference to a previously used configuration even when another 
configuration might offer some advantages. 

Let us give an exanple to illustrate the point. The pay load of a helicopter 

43 



with standard rarige amoiints to about 20^ of flying weight. Depending on the 
helicopter design, this weight may vary by 5^ to either side. This means that, 
at identical takeoff weight, a good helicopter will lift a payload weighing 25% 
of the machine itself, whereas a poorly designed helicopter will lift only 15^ 
of its weight. Thus, the second helicopter will be quite inferior with respect 
to lift capacity. This should be taken into account, on the one hand, in solving 
the problem of selecting the helicopter configuration and, on the other hand, 
in estimating the rationality of some particular configuration on the basis of 
a conparative analysis of data of escLsting designs. 



kh 



CHAPTER II 2M 

ROTOR AERODyWAMICS 

Section 1. Development of Rotor Th eory and Methods of 

Eroerimental Determination of it s Characteristics 

The idea of using a rotor in place of a wing as a lifting system was born 
in 1923' The Spanish engineer Juan de la Cierva, after the airplane of his 
design had stalled and crashed, decided to develop an autorotating sustaining 
system whose wing-blade combination would not lose speed at a low or even zero 
forward speed of the apparatus . 

An experiment carried out by him in 1924 in the laboratory of Quadro- 
Ventos in Madrid, which showed the lonlikely high values of the aerodynamic effi- 
ciency of an autorotating rotor as a lifting system. Induced a theoretical in- 
vestigation of the aerodsmamics of autorotating rotors, carried out in England 
by Glauert in I926 and later developed in 1928 by Lock for the case of hinged 
blades . 

Thus, experiment generates theoiy and the endeavor to extrapolate the re- 
sults of theoiy to practice gives rise to new experiments which more thoroughly 
reveal the pl^sical nature of various phenomena, which in turn leads to a new 
close examination and development of theory. Only in the unity of theory and 
practice is it possible to describe the development of rotor aerodynamics vp to 
the present state of the art. Thus, in 1928 the Glauert- Lock theoiy was first 
published. As is known, in this theory the magnitudes of thrust T, longitudinal 
force H, and torque M^ are determined as a function of the kinematic parameter 

V cos a 

of the angle of rotor blade setting cp, and of the flow coefficient 

V sin o — V 



A: 



"R 



which represents the ratio of the velocity of the air flowing through the disk 
to the tip speed tuR (Fig. 2.1). Consequently, the coefficients of the moments 
and forces can be expressed as 






(1.1) 



According to the momenttmn. theorem (see Sect .3), we have /l43 

45 



After determining from this the induced velocity v 

^ ____CrQJt/?'lo)2 _ Ct-u/? 



v = 



2Qn/?2U' 4QIt/?3(0 y^|i.2+ X2 4 )^(i.2+ X2 



(1.3) 



and substituting this expression into the formula for \, 



), = jJ, tan a - 



4 lA^2+ X2 ' 



we obtain the e^qaression for the angle of attack of the craft: 



tana = 



4(1 y |x-'+:):2 



(1.4) 



For calculating the aerodynamic characteristics of an autogiro. Lock pro- 
posed deriving the unknown value of the quantity \ from the conditions of auto- 
rotation, i.e., equating the ex- 
pression for torque on the rotor 
shaft to zero: 




V cos of: „., ./, 




4ufogi''° 



Fig. 2.1 Velocity Conponents of Air 
Flowing through the Rotor Disk. 



yW^=/((p,),(x) = 0. 



(1.5) 



Since the angle of blade set- 
ting cp and the characteristic of 
the regime |j. were prescribed 
quantities, the value of \ was 
determined from the quadratic 
equation (2.47) representing 
eq.(1.5) in a developed form. For 
a given value of the velocity of 
rotation uuR, flying speed V, 
angle of setting cp, and known \, 
the only possible angle of attack 
of the autogiro a at which steady 
autorotation occurred was found 
from eq.(1.4). In this case, the 
equation (see Sect. 2) 



^ = ^=/('f, !J-,/v) = a„ 



[\H 



3 ' 2 



(1.6) 



yielded the value of the thrust coefficient while the corresponding equation 
furnished the value of the coefficient of longitudinal force so that the polar /49 
of a freely autorotating rotor could be detemdned as a function of the angle of 
attack. This polar, Hke the polar of a wing, could be used in the aerodynamic 



46 



calculation of the autogiro. Such a method was published in 1931 by I.P.Bratu- 
khin (Ref.ll). 

In 1934, the author"" (Ref .4), in his study of overspeed of the autogiro 
rotor during the takeoff run, determined the unknown values of thrust and torque 
from eqs.(l.4), (1«5)» and (1.6) and prescribed values for the angle of attack ot , 
angle of blade setting 9, and initial peripheral velocity and flying speed, 
i.e., the parameter |j,, which made it possible to calculate the aerodynamic char- 
acteristics for cases of unsteady motion. The following method of determining 
the flow coefficient \ was proposed: 

Substituting expression (1.6) into eq.(l.4) will yield a polynomial of the 
fourth power in both X and |j,; thus, it is difficult in practical work to deter- 
mine \ at known ij, or vice versa, since each time it would be necessary to solve 
a fourth-degree equation. A study of the dependence of X on |j, at given values 
of a and cp shows that, with the exception of a small segment of negative X near 
|j, = 0, the curve \ = f(p.) represents a straight line with a high degree of accu- 
racy. The equation of the family of these straight lines at different values 
of Q- and cp has the form 

n~"-- (1-7) 

Equation (1.7) yieHs the dependence ofX oniJ,,Qf,9, and a; its use, to- 
gether mth the above formulas, permits calculating the thrust and torque coef- 
ficients. It is thus possible to calculate the aerodynamic characteristics of 
the autogiro rotor for any unsteady operating conditions when the torque created 
on the rotor by the air flow produces spinning or braking of the blades, depend- 
ing on the angle of attack and the value of |j, . 

This still left the necessity of finding a method of applying this theory 
(developed by Glauert and Lock for an autogiro) to calculation of the aerody- 
namic characteristics of a helicopter, i.e., to their determination under forced 
rotation of the rotor by an engine. Such a method was proposed by the author 
in 1945, in collaboration with V.N.Yaroshenko (Ref .9). 

To determine the torque necessary for fUght under given conditions it is 
logical to use the above system of equations. Here, it was convenient to pre- 
scribe the value of the thrust coefficient t, since the rotor thrust is easily 
determined if it is approximately assimied that the thrust, under steady hori- 
zontal flying conditions, is equal to the helicopter weight. 

Using prescribed values of the rpm and determining t from the expression 

^= o ^ • (1-8) 

2 



^ Here and elsewhere in Section 1, the author is M.L.Mil' . 

47 



the quaxLtity X can be found from eq.(l.6), the torque from eq.(2.47) and the 
angle of attack of the helicopter a which corresponds to these conditions, from 
eq.(1.4)» 

Even before the present stage of helicopter development, practical auto- 
giro engineering required the solution of certain stability problems; designers 
in various coimtries attacked the problem of determining the damping produced 
by the hinged rotor d\3ring vibrations of the craft. Analyses of flight accidents 
■with autogiros showed the necessity for studying the flapping motion of blades /50 
and for finding methods of retaining autorotation during aerobatic maneuvers of 
the craft . This led to the theory of a rotor for hinged blade attachment with 
curvilinear motion, which the author developed in 1939 and which represents a 
more general case than the Glauert-Lock theory established for steady recti- 
linear motion. 

Finally, in 1940 A .N JVIikhaylov worked out a method of equivalent rotors, 
which sinplifies the application of the Lock theory to a rotor equipped with an 
automatic pitch control mechanism (Ref.l5). 

The application of these methods to the aerodynamic calculation of Soviet 
autogiros between 1931 and 1940 and of the Mi-1 and Mi-4 helicopters in 1947-1952 
showed highly satisfactory agreement between design characteristics and flight- 
test characteristics . Since these first autogiros and helicopters flew at rela- 
tively low speeds and thus at small values of |j,, the inaccuracies of the theory 
due to the assunption of smallness of this parameter, made by Glauert and Lock, 
were nonessential. However, more coiiplex problems were still to come: the de- 
velopment of more powerful and faster helicopters, which inplied constant im^- 
provement of the theory. 

The Glauert-Iock theory, as is known, makes a number of assunptions (in- 
cluding uniform distribution of induced velocities) so as to permit integrating 
the equations in a finite form. Thus, the Cy of the section was expressed as a 
linear fionction of the angle of attack Cy = a^^ a, while Cxp was taken as some 
average quantity independent of the angle of attack. The forces acting on the 
profile, i.e., on the section of the disk where - in forward flight - the air 
flows around the blade from the trailing edge (this region is small at small 
values of |j.) were inaccurately determined. The radial conponent of the resultant 
velocity in the blade direction was also disregarded. The blade itself was 
assumed to be rectilinear, flat (not twisted), and of constant chord. 

During the period of 1932-1943, many researchers - Whittle and Bailey in 
the USA, Hohenemzer and Zissing in Germany, and others - further refined this 
theory in that methods were foimd for integrating equations in a finite form 
while doing away with many previously accepted assutiptions . The concept of the 
effective radius of a blade, smaller than the actual, was introduced for taking 
account of tip losses. The coefficients Cy and c^ represented more conplex 
functions of the angle of attack, etc. 

The most inportant inprovement of the classical theory during the postwar 
years was the application of methods of n^umerical integration to the calculation 
of flapping motion and aerodynamic forces. This permitted the direct use of the 
e^qjerimental characteristics of profiles, taken for the necessary value of the 

48 



Reynolds nimbers and Mach mjmber, in deterniining Cy and Cx of the section as a 
function of the angle of attack and thus to take into account also the effect 
of conpressibility. 

Later, it became possible to introduce into the calculations not only the 
initial geometric shape of the blade but also its deformations due to bending 
in the plane of thrust and in plane of rotation and, what is especially im- 
portant, due to torsion. 

However, the use of such a cumbersome method for practical calculations 
became possible only after appearance of electronic conjjuters. Thus, the modern 
method of rotor calculation, as presented in this book on the basis of studies 
made by A.S.Braveiman and M.N.Tishchenko, was a step-by-step process. 

However, even after all these refinements there still remained the /51 

rather rough assunption of a uni- 



Section where measurements 
made 
1 




Fig. 2. 2 Induced Velocity Distribution 
in Hovering. 



form distribution of induced ve- 
locities over the rotor disk, 
which not only led to an inaccu- 
racy in determining induced power 
losses but also to errors in de- 
termining the true angles of 
attack of individual blade sec- 
tions and hence to errors in the 
profile power, thrust, and longi- 
tudinal force. 

Thus, further refinement of 
the theory could be expected from 
the development of the vortex 
theory which is the only one 
capable of determining the distri- 
bution of induced velocities in 
relation to the forces acting on 
each given blade element. 

However, development of such 
a theory required greater insight 
into the plgrsical aspects of the 
phenomenon. Here again it was 
necessary to resort to e^qjeriment. 



Experimental studies of flow 
around a rotor in a wind tunnel 
followed by removal of the induced velocity field, carried out in 1946 \s^ the 
author together with M.K.Speranskiy, clearily showed that the vortex system known 
from a propeller operating under conditions of axial circulation flow and repre- 
senting [for the case of circulation constant along the blade (F = const)] a 
central vortex with blade-tip vortices shed by the blades, is transformed - at 
small values of p, - into a system similar to the rectangula.r vortex system 
characteristic for a wing. In turn, the induced velocity distribution obtadLned 
from experiment (Pigs. 2. 2, 2.3> and 2.4) fully confirmed the possibility of an 



49 



0.1S 



Tr 


"P 


ezoida 


bladc 






A- 


az u=-5' 




a: 



i°^ 



rn/sec R 

f 






\0.S 



-I 



s 



as 



V 



^=0 



as 



-1 



-2 



0.5 



I 



Z^ 



Fig. 2. 3 Induced Velocity Distribution in 
Forward Flight (|j, = 0.2). 



approximate representation of the induced velocity field, proposed by Kusner and 
Glauert, in the form of a funnel during the hovering phase (Fig. 2. 5) and in the 
form of a cylinder trimcated iDy an inclined plane during forward flight 
(Fig. 2. 6). This configuration could be used for refining the Glauert-Iock 
theory Td.thout resorting to the vortex theory. These e^qDerimental facts gave 
rise to a sequence of theoretical works, which were based both on the approxi- 
mate vortex pattern of rectangular vortices suitable for describing conditions 

at medium and high flying speeds 
and on various more general 
theories which examine a system 
of vortices trailing from each 
element of the blade. 

It is necessary to say that 
even before appearance of these 
ejgjerimental data, G.I.Maykopar 
examined the vortex theory of 
rotors, having proposed that the 
vortex cylinder slopes in direc- 
tion of flight, which then 
served as incentive for a series 
of more or less accurate studies 
in this area. 




Fig. 2. 4 Three-Dimensional Model of Induced 
Velocity Field of a Rotor. 



Subsequently, L.S.Vil'dgrube 
was successful in developing the 
vortex theory in the USSR, fol- 
lowed later by V.E.Baskin. 



50 



In all these theories for determining the angle of attack of the TDlade /53 
section it was necessary to define not only the known kinematic parameters o?, p., 
i|r, and 9 but also the vertical coirponent of the induced velocity v as a result of 
the action at a given point of all vortices in the region surrounding the rotor. 
The induced velocities are found isy the Biot-Savart formula as a function of the 
circidation in the blade section: 

Hence it is clear that, since the induced velocity determining the angle of 
attack of the section is a f\inction of circulation and since the latter, in 
turn, can be determined in terms of Cy which is a function of the angle of attack 
of the section in which the induced velocity enters, all these problems reduce 
to quite conplex ecfuations and can be solved only by approximation methods and, 
in particular, by methods of successive approximations . Generally, first the 
magnitude of Cy is determined under the assimption of constancy in the distribu- 
tion of induced velocities over the disk, after which the calculation is re- 
peated during which process the induced velocity is determined from the vortex 
theory. The integral value of the coefficient of thrust is then conpared with 
the assigned value, and leads to calculation of successive approximations. 



Vr'Kr- 




V COS ct 



Plane of rotation 




^(^,f) = v+arcasfj 



Fig. 2. 5 Induced Velocity Distribution Fig. 2. 6 linear Induced Velocity 
according to the Law of the Triangle Distribution. 



Without here examining the essence of various vortex theories which, spe- 
cifically, differ by the assunption of finiteness or infiniteness of the number 
of blades, or the principles of methods for calculating the induced velocity 
(see Chapt.II, Sect.5), we will estimate their general significance and role in 
the refinement of calctilations of the flying characteristics of the single-rotor 
helicopter. 

Let us examine the maximum value of the air in determining the induced 
power, assuming a unifonn and funnel-shaped induced velocity distribution. 
After this, we will conpare the induced rotor power under the condition of con- 
stancy of thrust for the case of a uniform induced velocity distribution over 
the disk and corrections for the funnel. 

The funnel-shaped induced velocity distribution is represented by the law 

51 



of the triangle (Fig. 2. 5): 

where v^ is the induced velocity at a radius r of the blade. 

According to the momentum theorem, the elementary thrust is /54 

dT=Q2nr2v'i. 
Integrating from to R, we obtain 

where v^ is the induced velocity in the blcLde tip section. 

For the case of uniform induced velocity distribution, the thrust is 

7",=q2jt/?2z,2_ 
Equating Ti and T, we obtain 

Qn/?2t)2=2jr/?2Qx;2, 

whence 

v^ = vy2. 

Consequently, 

v, = v^ —=.v— y'2. 

The induced power is determined from the following formulas: 

dN-^„^ =dTv = AnQv\rdr\ 



R 



Substituting into the last formula the value of Vj. = kr, we obtain the ex- 
pression for power in the case of an induced velocity distribution according to 
the law of the triangle: 

A^i„4 =— /?2Q23/2'r;3=2.26n/?2Qx;3. 

tri 5 

This power is greater by a factor of 1.13 than the power Nj^^ for the case 
of -uniform velocity distribution, since 



52 



Thus, the funnel-shaped induced velocity distribution increases the induced 
losses by about 13%' Consequently, the difference in the relative efficiency 
amounts to about 10^. This is a high value so that, for calculating the thrust 
and power of a rotor in a hovering regime, it is logical to use methods of the 
momentiom or vortex theoiy which permit talcing into account the additional power 
loss due to nonuniform induced velocity distribution over the disk. 

In an approximate calculation, the induced losses calculated by the formula 

can be increased by about 13%. 

In hovering, the expended power is the sum of induced power and profile 
power, where the former amounts to 75^ of the total required power. At cruising 
speed, the induced power is only 20 - 30% and at maximum speed, about 10^. Thus, 
at high flying speeds the maximim refinement in the required power, as a result 
of taking the nonuniform induced velocity distribution into consideration, /55 
can be not more than 1 - 2^ in power and not more than 1% in flying speed. 

Of course, using (analogous to the effective aspect ratio of an airplane) 
an effective rotor radius somewhat smaller than the actual radius, it becomes 
possible to introduce an experimental correction to the aerodynamic calculation 
which had been based on the theory of constant induced velocity distribution. 
This correction may be particularly inportant in determining the rate of climb 
and ceiling for heavily loaded rotors. 

Thus, it is obvious that, in calculating the flight data, there is no need 
for much more conplex calculations of the aerodynamic rotor characteristics 
based on the vortex theory. 

Refinement of the section angles of attack given by the vortex theoiy be- 
comes necessary only in calculating the stresses set vp in the blade, especially 
at low speeds where the aerodynamic forces inducing blade vibrations in second 
and higher harmonics actually are the result of the blade encoixntering the 
vortex field. Accoixiing to the theories stipulating v = const, these stresses 
- for all practical purposes - are equal to zero. 

It must be assimied that further refinement of the velocity field may be of 
importance in determining the boundary conditions at which flow separation be- 
gins. However, for this purpose it suffices to refine only the pattern of the 
induced velocity distribution in the form of longitudinal pitch, which leads to 
a variation in the flapping motion (bi) and to some redistribution of the angles 
of attack. 

At the same time, the concept of induced velocity distribution in the form 
of a funnel (Fig. 2. 5) and especially the assunption of its b-uildiqp from front 
to rear (Fig. 2. 6) result in substantial variations in the flapping motion (spe- 
cifically in bi ) and in the lateral force S, which has been taken into accoiont 
by various authors. 



53 



1. Classification of Rotor The ories 

It is obvious from this account that only two of the theories of a real 
rotor differ fundamentally. Both are based on a study of forces acting on the 
blade element. Such an approach to the helicopter rotor was first used by 
N.Ye.Zhukovskiy (Ref .1) as long ago as in his investigations on the effect of 
wind on the thrust of a helicopter rotor. 

In the first of these theories, the induced velocity distribution over the 
disk is prescribed, regardless of the forces acting on the blade elements. Its 
average value can be determined from the moment lan theorem. 

In the second theory, the induced velocities of each blade element are a 
function of forces acting on all blades, which in turn are a function of these 
induced velocities and are usually determined by means of the Biot-Savart 
formula. 

Let us call the first the classical theory which enconpasses the Glauert- 
Lock theory and its subsequent development, while the second will be designated 
as the vortex theory. 

The classical theory is conceivable with integration in finite form, while 
the vortex theory proposes only numerical solution methods. 

It is also of use to study the momentum theory of an ideal rotor, which can 
be used in developing energy methods for aerodynamic calculation and in inter- 
preting the results of an experimental determination of the aerodynamic rotor 
characteristics . 

Although the development of the classical theoiy in its numerical methods 
is nearing conpletion due to the use of conputers, the vortex theory still /56 
presents various problems. Thus, in most vortex theories, when calculating the 
circulation flow, the quantity Cy is taken as a linear function of the angle of 
attack since it can be refined. For sicplification, the vortex system is con- 
sidered to be two- or three-dimensional but linear. This sinplification can be 
conpensated later. 

Thus, the vortex theory, having inherited all refinements introduced dioring 
the development of the numerical calculation methods of the classical theory, is 
presently becoming the most accurate theory. In its development, there is no 
need to use the assunption of steady flow around the blade sections and the 
section polar can be refined Isy using ejgjerimental data as to the influence of 
centrifugal forces on phenomena in the boundary layer. 

Proceeding from this classification, let us give a further account of the 
rotor theory. 

2. Development of Experimental Methods 

Ibiperimental methods for determining the aerodynamic rotor characteristics 
are being developed simultaneously with the described development trends of the 

54 



r 



theoiy. 

After the first experiments in Madrid in the 1930s, researchers in mar^r 
coiintries began experimenting vdth rotors in wind tunnels. Experiments based 
on procedures by V.G.Petrunin at TsAGI were carried out (1931-1936) on models 
of autorotating rotors of 1.2 m diameter, in which the three conponents of force 
and torque were measured and the blade flapping recorded. 

The experimental technique was greatly inproved. In particular, it became 
possible to obtain the fuselage polar in the presence of an operating rotor. 
Measurement of flapping at the hub of an autorotating rotor, tiurning in the 
tunnel air stream about the translational velocity vector [which permitted esti- 
mating the rotor danping in roll (see Sect. 2)] was the most noteworthy achieve- 
ment by V.G.Petrunin in these e35)eriments . 

However, it soon became clear that small Reynolds numbers in the case of 
flow around the blade section led to such extensive distortions of the profile 
polar on the model, in conparison with a full-scale craft, that it was impos- 
sible to use these results directly. 

In 1944, the author together with I.F.Morozov at TsAGI set up experiments 
on a rotor of D = 2.5 m, which no longer concerned only the autorotation regime 
but also included the helicopter regimes . 

The testing facility for the rotor model of 2-5 m diameter is shown in 
Fig. 2. 7. Tests on this model were first made in the coordinates 6, mt = f(cp) 

for Li = const and t = const (see Fig. 2. 8), where 6 « ,^ . The tests, conducted 

for three types of blades of different shape and twist, made it possible to 
judge the propulsive properties of the rotors. However, the main thought behind 
these experiments lay in the possibility of estimating the degree of perfection 
of the theory in conparison with calculations. 




Fig. 2. 7 Testing Facility for Rotor Models. 



55 



I mil II iiiiiii 11 1 mil 1 1 



W. 



n f 


-^ 


p^ 


~ 


[ — 


p- 




.0 






^: 


^? 


- 


— 




- 


— 




— 


— 




— 







































A 
"^ 


•-" 



i2,!2f 



S./ff 



0.15 



DlOcp 




Whereas our first helicopters were calculated with reasonable acciaracy by 
the classical theory presented in Section 2, the Mi-6 helicopter vrLth lljOOO-hp 
turboprop engines had already passed into a region of flying speeds - and hence 
of values of [j. (more than 0.4) and Mach numbers - where the Glauert-Lock theory 
yielded appreciable errors. In preparation of this, a full-scale helicopter 
test stand was constructed at the TsAGI on which the author, together with M.K. 
Speranskiy, determined the characteristics of l4.5-m rotors in a large wind /57 

timnel. The next problem was that of 
converting the e^qDerimental data ob- 
tained for a specific rotor to another 
load factor and to other Mach numbers. 
These calculations are presented in 
Section 6. 

Because of the need to refine the 
theoretical methods of calculating 
aerodynamic loads on a blade element, 
a series of experiments were carried 
out in recent years, whose results 
jrield a more detailed explanation of 
the circulation flow pattern around 
the blade profile under actual rotor 
operating conditions. In this respect, 
interesting data can be obtained by 
measuring the pressure distribution 
over the blade chord in flight. By /5S 
integrating the pertaining pressure 
distribution diagrams, it is possible 
to obtain the actual inr-flight aero- 
dynamic loads acting on the profile 
and varying relative to the rotor azi- 
muth. In making such experiments on 
models, V.E.Baskin and A. S.D Dyachenko 
found that, as the blade passes into the region of vortex trailing from the tip 
of the advancing blade, there is a marked junp in blade loading. 

No doubt, the further development in helicopter engineering toward an in^ 
crease of fl3ring speeds will require new experiments and the development of test 
stands for experimental determination of the aerodynamic characteristics at even 
larger values of (j, while retaining similarity of Mach and Reynolds numbers. 

Section 2. Classical Theory of a Ro tor with Hinged Blade 
Attachment; General Cci.a eL Curvilinear Motion 

The rotor theory for rectilinear motion of a helicopter or autogiro was de- 
veloped by Glauert and Lock and has been described at numerous occasions [see 
(Ref.36, 37, 2, 11, 23)]. The theory presented below is a further development 
of the Glauert-Lock theory, for the more general case where the helicopter is in 
ciirvilinear motion so that the rotor axis describes a rotary motion in space. 
Such motion takes place during steady ciirvilinear flight of the craft, for ex- 
anple, during turning and also during vibrations relative to the longitudinal or 



-o-t-0.1; -a- t=- 0.125 -Trapezoidal blade 
— t-O.I; •^■t - 0.125 -Rectangular blade 

Fig. 2. 8 Aerodynamic Rotor Charac- 
teristics (Model Test). 



56 



transverse axis caused isy piloting or external factors. 

Rotor Theory in Curvilinear Motion 

Here, all the assxmptions made by Lock are accepted in the general theoiy, 
namely: The induced velocity in the absence of rotation of the rotor axis is 
considered as uniformly distributed over the rotor disk; the change in the law 
of its distribution over the disk pertains only to the siperposed effect of ro- 
tation of the entire craft, i.e., the initial pattern remains the same as for 
lock^s pattern. 

likewise, we adopt the assun^jtions of linearity Cy = f(ff) and the admissi- 
bility of replacing the coefficient Cx by some average quantity c^p which is 
identical for all blade sections. 

We will take into account blade tip losses, i.e., consider that no lift is 
developed at some tip portion of the blade and that the drag forces, just as the 
inertia forces, act on the entire radius. The method of obtaining the expres- 
sions for forces and moments here is analogous to the Lock method, so that we 
will not repeat the derivation of the fundamental equations here and only dis- 
cuss expressions that differ in the case of curvilinear motion. 

A special role is played here by the Coriolis forces of inertia arising 
upon rotation of the rotor. These forces will be discussed in most detail. 

The resiolts to be presented here contain the basic relations of the 
Glauert-Lock theory for rectilinear motion and, furthermore, permit answering 
the following questions: 

a) What is the flapping motion of the blades in curvilinear motion of 
the entire craft? 

b) How does the position of the aerodynamic resultant change in the case 
of rotation of the entire craft in some direction (for exanple, to 
the left or to the right) and is there a tendency to accelerate or 
decelerate this rotation? 

c) What peculiarities does the autorotation regime of a rotor exhibit /59 
in the presence of rotation? 

Along with an investigation of these problems we will also discuss the ef- 
fect of the configuration of the hub, profile, and blade centering on the be- 
havior of the rotor from the viewpoint of stability and safety of the craft. 

The obtained resiilts are common for any rotor with hinged blade attachment, 
be it the rotor of an autogiro or of a helicopter. 

In addition to answering the above questions which are of independent in- 
terest, the res\J.ts of the analysis yield some necessary data for studj^ing the 
controllability and dynamic stability of the mentioned craft. 

1. Coordinate System and Physical Scheme of the Phenomenon 

Coordinate system . The phenomenon is examined in a coordinate system fixed 

57 



vd-th respect to the craft (Fig.2.9)« This is a right-handed system (for right- 
hand rotation of the rotor) in which the z-axis is directed along the axis of 
rotation of the rotor and the x-axis backward with respect to the direction of 
the velocity of the center of gravity of the craft. The angular position of the 
blade ;jr is reckoned from the x-axis. 

Physical scheme of the phenomenon . We will study a rotor rotating in space 
together with the craft at a constant angular velocity and, in so doing, main^ 
taining a constant angle of attack with the fUght path. 

Such rotation occurs, for exanple, in turning. 





Fig. 2. 9 Coordinate ^stem. 



Fig. 2. 10 Velocity Diagram of Rotor 
in Turning. 



We are interested here in the manner in which the radius of rotation of the 
craft is to be taken into consideration. 

For this, let us examine a helicopter executing a turn of radius p at a 
large angle of bank Yt > close to 90° . 

The velocity diagram in the plane of symmetry of the craft, in this case 
close to the turning plane, is shown in Fig. 2. 10. 

The relative flow velocity at a distance r from the center of the rotor can 
be divided into two conponents: velocity Qp directed along the tangent to the 
path of the center of the rotor hub (we will call this the tiu"ning speed) and 
velocity ^r directed perpendictilar to the rotor plane. 

The distance from the center of the rotor hub to the center of gravity of 
the helicopter will be neglected for the radius of ctirvature of the path. 

If we now transfer the center of rotation of the craft to the center of /60 
the rotor hub, it should become possible, in addition to the indicated veloci- 
ties, to also allow for the additional centrifugal forces acting on the blades. 



58 



As is known, in the case of a steady turn without sideslip, the resultant 
of the centrifugal force and the force of gravity lies in the plane of syiimetiy 
of the craft. 

Thus, these forces, proportional to C?p, can be regarded as an increase in 
blade weight. Obviously, this increase is the overload n. 

The expression for the coning angle ao contains a term taking the blade 
weight into account: 



(2.1) 

where S^.i, is the static moment of the blade weight relative to the horizontal 
hinge. 



For turning, we thus have 



^%- /,-.V (2.2) 



It is known that, in turning, the revolutions of the rotor increase. If 
(Dt = ('u/rT, which roughly takes place during autorotation, then, substituting cut 
into eq. (2.2), we obtain 

Atto^Aao^, (2.3) 

i.e., dioring steady curvilinear motion of the rotor without sideslip, the de- 
crease in coning angle due to the weight of the blades remains constant regard- 
less of the overload. 

In the usual case, Aao is not more than 0.3°; this quantity can be either 
neglected or taken into accoimt, regardless of how this is done in the theory 
for rectilinear motion. Therefore, we will here discuss the following scheme: 
The rotor moves at a constant angle of attack and executes a rotary motion rela- 
tive to the axis going through its center. 

2 . Ine rtia Forces Acting on the Blade 

At curvilinear motion of the helicopter, the rotor blade executes four 
rotary motions in space. 

First, it rotates about the hub axis Oz with an angular velocity uu ; second- 
ly, it rotates together with the axis Oz in space with a velocity Q having the 
conponents 0^ ^^d Oy; finally, it vibrates relative to the axes of the flapping 
and drag hinges making an angle p = f (f ) with the plane perpendicular to the hub 
axis and an angle § = f(ilr) relative to its own mean position. 

Below, we will neglect the blade motion for the drag hinge, in view of its 
relative smallness . 

Let us examine the elementary inertia forces acting on the blade diiring 

59 



rotation of the rotor. The centrifugal force C 

dC=ma^rdr (2-4) 

is perpendicular to the rotor axis and directed along the radius . 

The centrifugal forces generated during rotation with angular velocities /6l 
Qx and Qy vd.ll be 

dCs^ = mQ'^^rs\nifdr, i 

dCsy=mQlrQos'^dr. \ (2.5) 

These forces also lie in the plane of rotation and, accordingly, are per- 
pendicular to the vectors Qx and Qy 

The inertia force of flapping, perpendicular to the blade, is 

dj^m^-^rdr. (2.6) 

The Coriolis inertia forces K produced by the rotations cu, Q^, and Qy, per- 
pendicular to the plane of rotation, are 

dK^^-= — 2ffi2^u)r cos <jj dr\ \ 

dK^ = — 2/re9y«)r sin <jj c?r. J (2.7) 

The Coriolis inertia forces generated during blade flapping, i.e., during 
rotation with an angiiLar velocity Q, are as follows: 

- conponent in the plane of rotation 

-^r dr\ 

(2.8) 



^ "^ " dt 

dKls=-2mQy^rdr; 
y ^ dt 



— conponent perpendicular to the plane of rotation 

(2.9) 



^ dt 



dKls =~2mQy<^-^rcos'!fdr. 

The Coriolis inertia forces due to flapping and rotation at angular velocity 

at/Cp„= -2/reo)B ^rdr (2.10) 

lie in the plane of rotation and are perpendicular to the blade. 
60 



The diagram of the inertia forces generated during curvilinear motion is 



given in Fig. 2. 11. The forces dC, dJ, dK'^Q 



and dKgQ are not shown. 



In studying the blade flapping relative to the flapping hinge, we will take 

into account only forces with respect 
to eqs.(2.4), (2.6), and (2.7), since 
any forces dKgQ [eq.(2.9)] of a higher 
order of smallness can be neglected 
for the inertia forces of flapping or 
for the forces dKQ and since also all 
centrifugal forces produced by rota- 
tions Ojc and Qy [eq.(2.5)] can be 
neglected for the principal centri- 
fugal force [eq. (2.4)]; the ratio 




Fig. 2. 11 Diagram of inertia Forces. 



-jjj-1 , for all practical purposes, is 
not greater than 0.01. 



These centrifugal forces /62 
Ceq.(2.5)] are accounted for in 
studies of the equilibrium relative to 
the axis of rotation in the e^qjression 
for torque. 



In addition to inertia forces, aerodynamic forces act on the blade element , 



With the aim of obtaining 




a clearer view over the effect of inertia forces 
and to check whether the general equations de- 
rived below correctly describe the phenomenon, 
we will first siitplify the problem. 

We will coii53letely disregard all aerody- 
namic effects and study the motion of blades 
hinged to the hub and rotating by inertia in a 
vacuum while their axis of rotation is turning. 

Motion of hinged rods rotating by inertia 
during hub rotation. Let us examine the general 
case of a hub whose horizontal (flapping) hinges 



are at 
tion. 



some distance t^ 



from the axis of rota- 



Fig. 2. 12 For Determining 

the Flapping Motion of the 

Rotor . 



Let the rotor rotate about the z-axis at an 
angular velocity ou (Fig. 2. 12). 



The axis of rotation z turns backward at a 
constant angular velocity Q, and, at some instant of time, is deflected through 
the angle 5 from the original position Zq. Taking, as reference plane, the plane 
perpendicular to the z-axis and assuming smallness of the angle 6, eqs.(2.4), 
(2.6), and (2.7) can be used for deriving the expressions for the moments of the 
acting forces relative to the axis of the flapping hinge. 



61 



The moment oi" the inertia forces of flapping (with the positive moment 
tending to raise the blade ijpward) reads 



^j=- ^m^ir-l^,J^dr = 



'kk 



f R R R 



"WW 



'h.h 'k.W 



Since, for modern helicopters, the ratio — ^'J' ■ usually is not more than /63 
0.02, we can assume that 



^ mr^dr^^mr-^dr^I^^; (2.11) 

R R 

j mrdr^^mrdr^S^^, (2.12) 



'w.y, 



and since the integral containing t^^^ can be disregarded, we have 



Mj--^{I^.^~2S^J^,^. 



Having designated 



u,, (2.13) 



we obtain 



'WW 

'w.w 



*'=-'- S-('-2.). (2.^) 



The moment of centrifi:igal forces is 

R 



M, 



(.= -Jp/ra<oV(r-/^J^r=-/v,.,«)2p(l-E). (2.15) 



Here, we assume the angle p to be small and consider that the vector of 
angular velocity w is directed along the axis of rotation z . 



The moment of the Coriolis inertia forces is 

R 

jM 



K9=-2jOTQyO)sln<J;r(r-/i,^,)fltr=-2/j,j,Qy«)(l-E)slni}>. (2.16) 



Equating the sum of all moments to zero, we obtain the equation of blade 
motion: 

^WW ^(1 -2^)+^v,.r=?(l -^)= -2/w.W V(l -^)sin'l'- 
62 



After dividing hj (1 - eT t neglecting e^ , and setting u) = -gl-, we finally 
obtain 

^-|-P(l + s)=-2^(l + s)sini.. (2.17) 

Let us find the particular solution of this equation. Setting p = N sin i)f 
and substituting into the equation, we obtain 

2Sv 
p= ^(l+e)sin.l). 



The general solution of this equation will be 



22 



p=/lcosl/l + e<i.+ Bsinl/l+e'l' (l+e)sin<j;. (2.18) 



Substituting the initial values of i|f = -^; P = 0; -^— = 0, we then /6k 

determine the values of A and B: 



2Qy , Jt 

£(0 2 

22y , It 

fi = ^^(l+s)sinVl-(-c— . 



(2.19) 



Substituting these values into eq.(2.18) and assuming ti = t ^ (here, 

■ill is the angle reckoned from the position of the blade at the initial moment, 
when t = O), we obtain 

f!^?^(cosl/r+lti-cos4',)(l+0- (2.20) 

£0) 

From this, it is easy to obtain also the solution of p for the case of 
hinges intersecting the rotor axis (l^.n - ^ = O) . 

We will evaluate the indeterminate form by the L* Hospital rule: 

22y — (cos/TT^h-cos+i) 2^ 
%^o= lim -•= 4'iSin'Jj,. 

Since the rotation is assumed to be uniform, then, having introduced i|fi = 
= cut, we find 

8 = 2< = ^^i, (2.21) 

and, substituting \|ri = i|f 2_, we obtain finally 

63 



P = 'ij COS i]) = 8 COS i|j . 



(2.22) 



This result can be obtained also by direct solution of the differential 
equation (2.17) in which it is assumed that e = 0. 

It follows from eq.(2.22) that, in the case of a hub with intersecting 
hinges with the rotor axis deflected backward at an angle 6, the blade will be 
deflected in the opposite direction through the same angle. 

This means that the plane of the blade tips in space remains constant while 
the rotor hub turns. 

The result is correct and has physical meaning. 

The solution for p is plotted in Fig. 2. 13 in the case of siperposed (e = O) 
and spaced hinges (for e = 0.2"; at -yj^ = 0.01. The curves were constructed 
on the basis of eqs.(2.20) and (2.22). 



/?i 1 1 1 1 1 [ 1 1 1 PT ' ~ • rrr "1 m ' '"T ur 


^^L " ' -^-^-ZiZ 




— " " ' 4 '1 


^^■f^f r ' /^ ' 


n -*"'""' l—\ <-» -■^ 1 


_-- — *^ '~ \ ] 7^1 


IT -'^^ f7i-\ 7^1 '~ /JV l: / \\ ' .: 


LL _^--?^. n\\ tlX y. ^ l\\ J \\ \ 


^-«J 1L\^^^. l^\' ~t ' -X - /. - 1 -, 1 \ l- 


n ^-^-"^ % 5 ^ ' ^^ r X. j^L _i i:^ .\. 


'' ■ 9r JM'm ist'isi JkUBJO 720 mi ^oo m m. '^ im '^wo^ moisjo mo mo im mrzoio ip . 


-^^^t/^ A 71 \Z, T i' \-^L \ L_L I. 


— -^--^ \ r i^r; \r/ \i ^^\ 


^-^-^^7 u, I:: \t;'l i^^t i 


-f? " -^ ^< ' - \i ^ ^v^ ^ 


— " ' - \\"^~--X^ 1 :5 ' v/ / J 






■ —— — Hinges auperposed -. . _^ ^^ i y 


■**■ 


^^ I J_ 'aL 



Ihl 



Fig. 2. 13 Flapping Angle p vs. Azimuthal Position of Blade. 



As indicated by Fig. 2. 13) over a certain interval of time the motion of the 
blades almost coincides, i.e., their position in space is constant; after this, 
the blade on a hub with spaced hinges gradually starts following the axis of /66 
rotation, i.e., the plane of rotation of the blade approaches the plane of rota- 
tion of the hub, with some phase shift. 



^ Such a large value of e is used to make the graph clearer. In practice, e 
does not exceed 0.02. 

64 



At this time, however, the angle of turn of the hub is already so great 
that a further analysis is useless since" the blades are pl:ysically restrained in 
their arresting devices on the hub and since the above theory holds only for 
small 6 . 

Thus, the study of short-time oscillations shows that the rotor with small 
hinge spacing is practically equivalent to a rotor with intersecting hinges, 
i.e., it retains its position in space unchanged. 





S?r 




sn 



Fig. 2. 14 Velocity Diagram of 
Blade Element . 



Fig. 2. 15 Change of Velocity Field 
during Rotor Rotation. 



In the subsequent analysis we will disregard hinge spacing, and assume 
e = 0. Furthermore, we have reason to believe that the aerodynamic forces (as 
will be shown below) entrain the plane of rotation of the blades beyond the 
motion of the hub so that equilibrium is established between the blades and the 
inertia forces that tend to keep the rotor position in space constant; thus, the 
plane of rotation follows the motion of the hub with some lag. 



3. Aerodynamic Forces Acting on the Blade 

Let us assume that the aerodynamic forces acting on a blade element lie in 
a plane perpendicular to the blade axis and depend only on the velocity com- 
ponent lying in this plane. 

The coordinate axes Ijning in a plane perpendicxolar to the blade axis are 
situated such that the Ox axis is parallel to the plane of rotation and the Oy 
axis, perpendicular to it, Hes in a plane containing the rotor axis. 

Let us now deconpose the velocity acting on the blade element into the comr- 
ponents U, and Uy (Fig. 2. Iff-) . These are equal to 



U, = XRm — r 






C/_f^r(o-i-(i/?(B sirup; 
- (ji^cBp cos <^ -\- Qyf cos (J) — QjT sin <]). 



(2.23) 
(2.24) 



Here, eq.(2.23) is the same as in rectilinear motion. In eq.(2.24) only 
the last two terms are added, which take into account the effect of rotary motion 
of the axis. Consideration of these terms gives the change in Velocity field of 
the rotor, which becomes analogous to that depicted in Fig. 2. 15- 



65 




.^ Taking the new values of U^ and Uy into 

dU^ account, all elementary forces acting on the 

rotor blade can be obtained. 



As derived by Lock, the elementary /67 
aerodynamic forces and moments are expressed 
in terms of velocity conponents of the ele- 
Fig.2.l6 Forces Acting on a ment in the following manner: 
Blade Element. 

dT =-i- M- (?f/^ + t/,f/,)rfr; (2.25) 

dM = ^bQ (c,^^^ Ul - a^^U^U, - a^Ul) rdr; ( 2 . 26 ) 

ar/y = ^^sin<j.-pflrrcos^; (2.2?) 

dS== ^cos(j> — pafrsintj^. (2.28) 

4. Equation of Moments ReIlative_to Flap:oing Hinge 

The equation of moments relative to the flapping hinge can be written as 
follows (see scheme in Fig. 2. 16): 

R 



— r ^ OTr2 oTr — r pOTu>V2 dr—^ 2mQyW sin ipr2 dr 



R BR R 

— f 2m2_fU)COs<Jjr2afr4- {dTr~ { mgrdr^O. 





R BR R V,2.27J 



After integration, we obtain 



BR 





let us change from differentiation with respect to t to differentiation 

dt 
with respect to i|f, assuming j|. = u) : 

dl_dl_ _ d^_d2l_ 2 
dt ~d']i ' dfi ~rf<j/2 

and then rewrite the equation of moments, after dividing it by I^.^uu^ : 



BR 

rV^~~ \ ^' ' '^ sill y ^ (-UOV . f ^ r^r^\ 

''•1^ ^h.K"0 " - /kK"2 (2.30) 










66 



Using eqs.(2.23) and (2.24), we write out the e^^jressions for U^ and U^Uy 
entering eq.(2.25) for dT: 

Ul = rW -f 2oy^Rr^ sin tj) -j- ^^R'^ay' sin^ i);; (2.31) 

i/^f/y =Xu)2^r — ^(or2 — p(i<o2/?r cos ijj — 2^u)r2 sin 4. + 

-f X(x(b2/?2sjha ^ !iu)/?r sin i]j— SaWy?^ sin il) cost!) — 

''^ (2.32) 

— 2_jU)(i/?r sin^ i[) -(- GyUir^ cos i{» 4" 2y«>(j./?r sin <(< cos >]). 

The thrust . moment is 

BR BR 



I B2 „ . „ , , B3 B4 rfp B3 B< Q^ , , , 

+ — cpp,2sin2,^+— X ^__p(j.cos<^-— --t sin(p + 

2 3 4 rfip 3 4 o) 

, B2 B2 B3 rfp . , (2.33) 

-| X[j. sm <jj ■ P|j.^ sin i}) cos ijj !- (a sini|) — 

B3 Qr B-l 2v . B3 Sv 



3 



— (J. sin2 J; -| cos <jj -| |j. sin dicos ij< ) . 

<i) 4 o) 3 to / 



boa R^ 
Let us substitute expression (2-33) into eq.(2.30) and set y = Jl "^ — ; 

this finally yields ^l^.h 

, B4 2y B'' 2y 1 B4 2 (2.34) 

H — ^ cos(;;+~ [j,sin2cjj-| -cp^ B^fu.sin'!j + 

4m 3io2 43 



-| (f\>.^sm^<l^)~2 — sint}) — 2 — cos'l' 

2 / M (O ^v, t, ' 



The particular solution of the differential equation (2.34) can be repre- 
sented as the series 

fi==ag — aiCOst|i — i'lSint}) — a2C0s2(j) — 62sin2i]) — . . . 

It is known from solutions obtained with retention of only five terms in 
the expression of p and from practical experiments that, in the usual case of 
(n = 0;, the second harmonics of the angle p are small in conparison with the 
first. 

For greater clarity and sinplicity of the derivations (while fundamentally 
retaining an acciu-ate pattern) we will discard, in solving the problem, the 
second harmonics in the expression for g, i.e., we take p in the form 

p=ao — aiC0S(;»-6iSin<|;. (2.35) 

67 



We then find the derivatives -gr- a^nd '^ g -- and substitute them into 
eq.(2.34). Discarding terms containing functions of double angles, we derive 

-[Mf-(5=+-f!^^)-f-«oYl^]cos.], = Y[f ^+f 5^fiH^^^)]- 
SLL , /B2 B4 Q r 2 D3 o ^y\ • , I (2.36) 
/^^(i)2\2 4<o o (1)/ 

Since eq.(2.36) is an identity, we equate the coefficients of sin ^ , /69 

cos i|f, and of the free term, whence we obtain the coefficients of flapping 
motion. 



The free term is 



„,.v[i|.+ Xsn.H,.')-«^.]-,^. (,.3,, 



The coefficient of sin i|f is 

B2 . 1 X B2 BiQ^ 2 Q„ 

from which it follows that 

1 /n.^^ , 8S: 



The coefficient of cos i|f is 



B2-^|.2 ^ •"' B2^B2_i-^2^ 



B'2 / 1 \ B3 



4<o 

whence 

4 B / Sv 82 

^=Yl^«o — (^*-f- ■ 



S'lQyY 2 ?:L 



1 \<o Vw// 1\ 

B2+— 1^ \ T^ /B2^62+— H.2J (2.39) 

For the conditions of rectilinear motion we obtain 



68 



a\ =2^{\^±. 5cpj 



K=^ V-°Q 



B2_^.2 



B2 + — ^2 



Then eqs.(2.37), (2.38), and (2.39) can be rewritten as 



I (0 yo) I 



6, = ^;-(fi-.^-«?i 

' <i) ya 



52(52—1-^,2) 



52(b2 + ^^2J 



(2.41) 



¥e have examined the particular solution of eq.(2.34). Investigations of 
the general solution of this equation characterizing the natural vibrations of 
the blade show that the blade is stable and, after having been deflected from 
its path normal for the given regime, will return to it under strong danping 
during one or two revolutions of the rotor. 



This means that the new flapping motion of the blades, described by 
eqs.(2.4l) and caused by the presence of angular velocity of rotation of the 
craft, is established rather pronptly (within one revolution of the rotor). 



/70 



The obtained expressions hold for rotors with large y. Equation (2-41) 
shows that, imder the effect of angiilar velocity, the cone of the blades rotates 
together with the shaft, lagging behind it by a constant angle Aa^ and is also 
deflected in a perpendicular direction. Thus, the plane of this new slope is 
shifted in phase relative to the plane of rotation of the rotor axis. 

This shift h^ in the case of longitudinal rotation is characterized by the 
relation 

tan Af]* = — !• , 

where Abi and Aa^ are determined by eqs.(2.4l). 

Equations (2.4l) can be used for conpiling a table (see Table 2.1) showing 
the direction of flapping of the cone under the effect of angular velocity. 

Deflection of the cone in the case of conplex motion of the rotor (for 
exanple, to the front and right) is easily obtained by means of eqs.(2.4l). 
Thus, for the particular case y = 8 and Q^ = Hy we obtain Table 2.2. 



69 



li 



TABLE 2.1 



TABLE 2.2 



Mode of Revolution 
of Rotor Axis 



Pitches 

Banks to the left 

Dives 

Banks to the 
right 



Mode of Deflec- 
tion of Cone 



Forward and left 

Right and 
forward 

Backward and 
right 

Left and back- 
ward 



Mode of Revolution 


Mode of Def 1 


of Rotor Axis 


tion of Con 


Dives and banks to 


Backward 


the right 




Dives and banks to 


To right 


the left 




Pitches and banks 


To left 


to the right 




Pitches and banks 


Forward 


to the left 





The deflection of the cone during rotation is easily determined if we re- 
call that it is deflected to the side opposite to the rotation (in the plane of 
rotation) and, furthermore, is deflected in a perpendicular direction opposite 
to the direction of the gyro reaction (a right-handed gyroscope, on sloping back- 
ward, is deflected to the right while the additional slope of the cone occurs to 
the left). 

5. Physical Meaning of the Obtained Result 

Equations (2.41) show that, in response to rotation of the rotor axis to 
some side, the axis of the cone in this direction will lag (the plane of the 
blade tips, relative to the shaft axis, is^ deflected to the opposite side) and, 
furthermore, will incline in a direction perpendicular to the rotation. 

It is not difficult to demonstrate that the lateral inclination is caused 
by aerodynamic forces and the lag by inertia forces. 

As a typical exanple, let us examine the case p, = 0; B = 1. 

Let us assume that the rotor revolves backward; Qy > 0; Q^ = 0. Then, /71 
according to eq.(2.24), 

t/y = X/?u) - r -^ - lj./?p cos '> + Qyr cos ,<p = Uy^^ + At/,, 



where Uy is the esqaression for Uy when n = 0. 

The angle of attack of the forward blade sections (i|r = rr), assuming \ = 
= const, decreases by 



A{/y 



while the angle of attack of the blade sections stationary with respect to the 
flow increases by the quantity 



70 



Under the effect of the change in moments of aerodynamic forces occurring 
in this case, the flapping motion of the blades mil change until a new equi- 
librium of the moments and a corresponding flapping motion are established. As 
follows from the condition of zero hinge moment, the angles of attack of the 
blade sections should retiirn to their previous value corresponding to recti- 
linear motion (since no new forces appeared and the velocities U^ remained 
constant). 

This is realized when the real axis of rotation of the cone swept by the 
blades is deflected leftward by the same angle Ao? (the axis is inclined to the 
right to reduce the angle of attack of the advancing blade in the case of regular 
stationary flight) . 

Actually, the formula for bi Ceq.(2.39)] will yield in this case: 

i.e., the slant of the cone to the right, occurring in rectilinear flight, de- 
creases by the indicated quantity. 

Now let us examine the effect of inertia forces. The additional force act- 
ing during the rotation is the force due to Coriolis acceleration. Its moment, 
accoiTding to eq.(2.7), is 



A1 



^= — f ^rnQyinr"^ slw!^ dr = — 2/^ ^ QyU sirn}». 



i.e., during backward rotation of the rotor the advancing blade ( \|i = -^r-] is 

acted on by the moment M^ = 2lh.h^yU^ which tends to depress the blade. 

This new moment changes the blade flapping. Equilibrium is established 
when, as a consequence of this flapping, the angles of attack of the blade sec- 
tions in a forward position increase so that the additional aerodynamic moment 
equalizes the moment of the Coriolis forces. 

An increase in the blade angle of attack at the position i|f = -— - is achieved 

when the real axis of rotation of the cone described by the blades is deflected 
forward (imder rectilinear flight conditions, the axis is inclined backward so 
that the angles of attack of the advancing blades diminish) . 

Actually, eq.(2.38) for ai yields Z22 

, 8Qy 

71 



i.e., during backward rotation of the rotor, the cone follows the rotor axis 

So 

with some lag, being deflected forward from it through the angle ^ 



yw 



6. Equation of Torque 



The elementary moment due to aerodynamic forces can be written in the form 
of eq.(2.26): 

dM^= 1- bQ {c.p^Ul - a^^U,U, - aMl) r dr. 

Here the braking moment is considered to be positive. In addition to these 
forces, additional inertia forces [eqs.(2.5), (2.8), and (2.10)] appear in the 
plane of rotation owing to rotation of the rotor axis in space. 

The expressions U^Uy and Uy in curvilinear motion can be represented ty 
eqs.(2.23) and (2.24) in the following form: 

t/fAd/y= — 9_^u)[),/?r sln^ i]j — Q^^ior^ sini}) -|- (2.L2) 

-j- Qycor^ cos i)j -|- 2yU)[j,/?r sin ijj cos ijj; 

^Uy = — Sj-r sin tji "1- 2yr cos <}); 

hUl = — 2Q^m\Rr sin ij; 4- 2 -?- Q^.wr^ sin <]> + 
-J- 2p9^u)ij./?r sin tj) cos I}) + 22yO)X/?r cos i|) — 
— 2 -^ Qy(or2 cos ij) — 2^QyW^Rr cos^ <]; — 
— 22^2yr2sin6cos<J) + 2>2sjn2^_j_Q2r2cos2<;>. (2.43) 

Then, transforming the expression dM,. and adding the moments due to inertia 
forces according to eqs.(2.5), (2.8), and (2.10), we obtain 



where 



dM^=dM^^^-\-AdM^ 

* ^st ^ inert ' 

AdM^ = _ -i- 6Qa„ i<fU^MJy + A^P rdr-\- 



-\-m 



.22yiisin^-22,^cos^+(Q^^- 2^) sin <p cos ^^ - ^2.44) 

~2^^.y^dr., 



72 



BR R 

t trf^J t«ro^J * inert (3.45) 



Assuming p = ao - ai cos i|( - bi sin i|r and substituting into eq.(2.44) the /73 
expressions U^AUy and AUp from eqs.C2.42) and (2.43), we obtain, according to 
eq.(2.45). 



BR 





— tfQyiur^ cos tjj — tp2ya)(j,/?/-2 s jntj, cos ([) -j- 22_jU)X/?H sin t}i — 

— 2a,2^<or3 sin^ i]j + 26iQ_f cor^sin tj> cos i]; — 2 (Oq — a, cos <]> — 

— ftj sin i]j) 2^iop,/?r2 sin ijj cos i]) — 2Qy (oX/^r^ cos tj) -|- 
-|- 2a,2y(Br^ simp cos <); — 2&i2yU)r^ cos^ tJ) + 

+ 2 (flg — a, cos tjj — di sin <J>) Q^coix/^r^ cos^ tji -j- 22^2yr' sin tjj cos tj) - 

R 

— 2>3sin2tl)— 2^r3cos2<|j]a(r+ f m[ — 2ai2yCosin2.j)-|- 



-j- 26,2yU) sin (J) cos tj* -[- (Qj/ — 2^) sin i}) cos ^ — 

— 2(ao — Oi cost]) — 6isintj()(ai sini]> — A; cos tjj) co^] r^ rfr. 



The moment due to z^ of the blades per revolution is 



^t = ^ J ^1 ^'l- = ^ t .t + ^ ^*ea„-R*'o^x 



X 




r53 Q_^ fi4 Q_^ 



r0» i.!^ i(< w^ ±(4Uy 

L D o) 4 a) 4u) 

+^--Te)"-T(^;]+ 



(2.46) 



Qj. Q« 






Here M^ is the moment due to aerodynamic forces, obtained in the theory 
for rectilinekr motion 

^t st= '"trt4"'^''*^^''"' = '"trt T""^ ^'"^^^ ^^' 

Ht ^lia^ ^ ^^ ' 3 ■ 2 8 I, ^ 2 '^ ; (2.47) 



B2 „ B3 ^2 / B2 \ ;,2 1 



73 



We then write the expression for nit ^7 means of eqs.(2-46) and (2.4?): 



r Cj-p 53 B2 

.2 



-^(i<'+Asv)-f»>=+f».V- 



+-r W 7 — «!--; f'i + -^ — ^oH'- 

D (1) 4 (d 4t»> 0(i> 



■ey-TCfy+T(v^-°v«.)■ 



7- Rotor Thrust and Angle p.f Attack /74 

The rotor thrust is 



Substituting the values of U^ and Uy from eqs.(2.23) and (2.24) and inte- 
grating, we obtain 

Setting 

T=l-toQnR*w-^ = ial~Q{wRyF, (2-49') 

we find 

L 2 ^ 3 V ^ 2 ^ ) 4« n (2.50) 

As we see, the expression for the thrust coefficient has been somewhat 
changed; nevertheless, assuming ^ = 0, we obtain the same expression as for 

the case of rectilinear motion. The formula for the angle of attack, after de- 
termining the induced velocity, is obtained from the expression 

r = 2jt/?2QV"i>, (2.51) 

where V is the resultant velocity, 

V' = Y{Vsina-vf + V?cos^a. (2-52) 

Substituting T and V by their expressions and remembering that 

74 



K sin a — v K cos a 

" V--- 



we obtain 

. X , at 

*" ~~~^'^ 4h./X2+7^' (2.53) 

i.e., the usual e^qDression for the angle of attack in which only t has a new 
value. Writing this in expanded form, we derive 



,_i,--[T-iKt")-^n 



,„«=--r. - ^^-=^ ■-. (2.54) 



8. lateral Force 



According to eq.(2.28), the elementary lateral force is 



We can represent dS as 



dS= icostjj — ^dTsin<^. 



ciS-=dS^^-\-^dS, 



where dS^t is the elementary lateral force for the case of rectilinear motion: 

AdS^-^^^cos'^-^AdTsiTK^. (2.55) 



In eq.(2.55) we can replace AdMt = AdM,.^ ^ since the inertia forces /75 
when summed with respect to z^^-blades will give zero. It is easy to demonstrate 
this by means of eq.(2.44): 






io^ COS <{<:;= — l™«£Lsin-j^ = 0. 

u 

Then, losing eqs.(2./i4), (2.42), and (2-43), we obtain 

ArfAft..„ J 

cos '> = — bQa„ [cp2_^(i)(j./?r sin^ i}» cos tjj -f 



r 2 

-|-tf£2_^u)r2sintJ;cosijj — cpSyfor^Cos^tp — tpQy(n[ji^rsin<l)cos^i}i-j- (^ rz'v 

-f 22^u)X/?r sin i); cos ([) — 2a,Q_f(Br'^ sin^.ij) cos 4> + 

-|- 26,a^a)r2 sin ^ cos^ t|) — 2(ao — a, cos ip — 6, sin t|j)Q^o)|x/?r sin <J( cos^ tjj — 

— 2Q^mKRr cos2 ij)-f 2ai2j,a)r2 sin <!/ cos^ <]; — 



75 



— 2&iSy(i)r2 cos^ ij> + 2 (Co — a, cos <!) — frj sin ^) Qy<B(j,/?r cos^ <!/ -i- 
+ 22^Qyr2 sin •]> cos^ <]; — 2 Jr^ sin^ <{) cos <!> — Q^r^ cos' <];] dr. 

Setting 

we find, by means of eq.(2.42), 

pA rfr sinif) = — bQa^ (a^ — a^ cos tj> — d^j sin ij;) ( — 2^o)|i/?r sin' <!> — 
— 2,(0^2 sin^ ijj + 2y0)r^ sin ip cos <]j + 2yO)|jL/?r sin^ (j> cos i})). 

The lateral force of the rotor can be represented in the form 

5=5,t+AS, 



where 

2<i B/? 



a5=-^ ?^cos<J-flfrar.j;- 



2r. BR 

~lrj f?Aa^rsin<J;flrrc?<|-. (2-59) 



(2.60) 



Substituting eqs.(2.56) and (2.5S) into eq.(2.59) and integrating, we 
obtain 

Substituting S^t ^7 its value of /26_ 

+^(B=X+^a.,.)]— |-aoi.(m+^cp)+ (2.6I) 

we finally obtain 

76 



+-^t(i— t*''^)+-;7t(t+'+t"'^)J- (2.62) 

9 . Longitudinal Force 

According to eq.(2.27), the elementary longitudinal force is 

or// =^^sin <); - p orr cos (|>. 
Setting 

we find 

AdMf 
^dH= — ^^sin.]>--PAafrcos<l;. (2.63) 

Analogous to eq.(2.56). 



*-oxro . 1 



•slni]) = — bqa^ 



(f)Q_j.o)[x/?r sin^ tjj -j- tpQ^mr^ sin^ <ji 



r ' 2 

— cQ^mr^ sin ijj cos <]> — cpQ„(u|j,/?r sin^ 4* cos (j) -\- 22_^u)X/?r sin^ li — 

— 2ai2_^o)r2 sin't}) -f 26iQ_jO)r2 sin^ <h cos (]j — 

— 2 (^o — a, cos (]< — ij sin ijj) 2^o:[j,/?r sin^ t); cos <J) — ( 2 .64) 

— 2QywKfir sin <]; cos iJj + 2a,2yU)r- sin^ ip cos tjj — 2bj^Qywr'^ sin iJj cos^ <{; + 
-j- 2 (ao — a, cos i]j — &i sin tjj) QyO)(j./?r sin tj; cos^ <^ + 

+ 2 Q_^2yr2 sin^ i|) cos tj^ — Q^r^ sin^ ;|j — Q^r^ sin ijj cos^ t{i] c?r. 

Analogous to eq.(2.58), 

pA ofr cos ijj= — - ftga™ (flg — a, cos <]j — Aj sin (jj) ( — 2^(fl|j,/?r sin^ ti cos 
— S^iur^ sin ij> cos ijj + Q^utr^ cos^ '^j + 2y(oiJ./?r sin ij) cos^ <])). 



(2.65) 



Let us use the same notations as in the calculation of the lateral force, 
namely 

2it BR 2ic BR 

^^^^^^ '—^'^^^^'■d^-f^fj^^drcos^drd^. ^^'^^^ 







Substituting here eqs.(2.64), (2-65) and integrating, we obtain /77 

77 



2 

B3_ 
6 



2a ^ ^=P>-|*+-V'P^i 



(2.68) 



The e25pression of the longitudinal force for rectilinear motion has the fonn. 

Summing eqs.(2.67) and (2.68), we finally obtain 

+T*+f-.^+f »>+°t ^iCf +^+T^-)- (,.69) 

-TT(°f+iv)]- 

10. Consideration of the Chan&e in the Law of Induced 
Velocity Distributipri du ri ng Curvilinear Motion 

In cujrvilinear motion, a change both in magnitude and in character of the 
induced velocity distribution over the rotor disk should occur owing to the 
presence of new forces, namely Coriolis inertia forces; the moments produced by 
these forces are balanced (in the case of hinged blade attachment) by the redis- 
tribution of aerodynamic forces. 

It is logical that, if the angles of attack of the blade sections at a 
given angular position in curvilinear motion do not change in con^iarison with 
those existing in rectilinear motion, then both the forces and the induced ve- 
locities caused by them also will remain constant . 

Assuming that the rotor revolves backward (fiy > O), the angle of attack of 
blade section, at iJ, = 0, B = 1, will be 

Uji ^/?w — rutoi sin ij; + rmbi cos iJ' 4- Syr cos <}' 

Substituting here the e^qjressions for ai and bi obtained from eqs. (2. 38) 
and (2.39) and setting p, = 0, B = 1, and Q^ = 0, i.e.. 



we obtain 



8Qj, 


^ = 


2y 







Y™ 




(1) 



X/? 82}, 

ar=<p-j s'm<]i. 

r •yo) 



The angles of attack in the forward and rear positions (^ =0, i|r = tt), as 
shown above, do not change when \ = const, i.e., the kinematic change in velocity 

78 



of the disk (flyV cos i|») does not produce a change in the angles of attack of the 
blades thanks to the corresponding change in flapping motion. 

The above statements also hold for p. ^ 0. The curves of the azimuthal /78 
change in angle of attack of the blade section, plotted in Fig. 2. 23, for recti- 
linear motion and for rotation of the axis in a longitudinal direction also show 
that, at azimuths t = and ijt = it, the angles of attack, for all practical 
purposes, remain constant*. 

The change in angles of attack, for the exanple under study, takes place 
from the right and left of this position, the maximum change occurring when 

^ = -S- and ^ = -i- TT, which corresponds to the azimuth of the maximum value of 

the Coriolis inertia forces. 

Thus, it can be assumed that the previous distribution law of forces and 
induced velocities over the disk is si?)erposed by aerodynamic forces equalizing 
the moments due to the Coriolis inertia forces and the resultant velocities. 
These forces have a maximimi in a plane perpendicular to the direction of rota- 
tion, so that the induced velocity field will be tilted in this direction. 

The vertical conponents of the Coriolis forces are ejqjressed, according to 
eq.(2.7)> in the following manner: 



R 1 

Ks = — ("2m2j,(orslniJjrfr= — 25j,^,Qv"'sintjj; 1 

R 

ATa = — f 2mQ^<f>r cos <\)dr = — 2S^^^Q^<» cos ip. 



(2.70) 



These forces are linearly distributed over the radius and are periodic. 

It can be assimied that the aerodynamic forces and their induced velocities, 
equalizing the Coriolis forces at each angular position, obey the same law of 
change both with respect to azimuth and radius of the blade. Then the induced 
velocity in the case of curvilinear motion can be expressed in the form 

i>==t)o + '"i-^sirnj.+t;2-^cos.l). (2-71) 

Here, the velocity directed downward is considered to be positive. Let us 
denote 

Py = =C 



vo T 

Px= =C ' 



(2.72) 



■«- The difference in the angle of attack of the section, at \|( = 270° , for rota- 
tion in a transverse direction (see Fig. 2. 22) is due to a change in X . 

79 



Here, I Kq I and I Kq I are the absolute values of the CorioHs forces at 

angular positions where they reach a maximum. In calculating Vq, it was assumed 
that this velocity is constant over the radius. Assuming an additional induced 
velocity as a linear function of the radius, we should, from the condition /79 
of equality of momentum, introduce some factor C into the esq^ressions for p^ 



•a 
and py . At C = —- 



Pr 



3Sv,.w^b 2y 



1 ' 
— aonRH 



(2.73) 



Substituting Vx and Vg ty their values from eqs.(2.72), we obtain 



R 



R 



We then e^qiress the mean induced velocity in terms of the magnitude of 
thrust. According to the law of momentum 



27t/?2 qV 



(2.74) 



Here V' is the resultant velocity determined from eq.(2.52). Substituting the 
expressions for T and V' from eqs.(2.49) arid (2.52), we obtain 



_ QtlS,R 



Let us denote 



X,= 



f 



it 



^R 4/X2 + H-2 
Then, the velocity conponent Uy of the blade element will have the form 



(2.75) 



f/' =X/?u) — /7 Xircoslntj)— p jXirw cos <]i~r — — 



dt 



■ |i,/?<Bp COS <{> + 2yr COS i]> — Qj^r sin i[). 



(2.76) 



To the expression U^Uy obtained for v = const, we then add the terms 

UjiA'Uy = — jOy^^ir^to^ sin (Jj — pyXj<fl2/?/-ix sin^ i)j — 
— pJ^irV cos ({j — Pj^i'o'^Rr^ sin i]j cos tj». 



(2.77) 



The expression for the rotor thrust will have the form 



80 



in BR 



r=^J J -|-&Qa-(9t/^ + t/,d/y + t/,A't/y)rfr-rf1', 







which, in the esspression for the coefficient t, will give the additional term 
In the expression for the thrust moment, we obtain the additional terms 

BR BR 



Substituting the expression for the thrust moment (with these additional /80 
terms) into eq.(2.30), we obtain the following expressions of the flapping coef-- 
ficients, with consideration of a variable law of induced velocity distribution: 



a, = 2,(x+-i- 5cp) L^ (fi^"l-+«-^+5V/,) 



B2+-i-^2 ^ - y^ ^B2(b2_J-^2)' (2.79) 



1 \ MVM / I \ 



Let us now derive the formula for torque in the case of a variable law of 
induced velocity distribution. 

We find the expression for Uy. Let us denote 

A'i/y = — p^irui sin (p — /t7^>-iru) cos <1>, ( 2 .81) 

where Uy is taken with respect to eq.(2.24), i.e., without consideration of a 
variable induced velocity. 

Then, , , o 

= U\ — 2kpykyR'r sin ^ — 2lp^^<i>'^Rr cos if + 

81 



+2 J^Py^yr^^in<^-\-2 ^p;kyr^ cos^+ 
+ 2P)PyXj{j,a)2/?r sin <j< cos <)» + 2^pj^^^m^Rr cos^ .ji _ 2Qympy\r'^ sin <p cos <}» + 
+ 22^a)Pj,Xir2 sin^ <]) — 2a,(i)yo^i/-2 cos^ <}. + 22^(j)/?^i sin tp cos <J> + 
+ /7^X2a)V2 sin2 .p + 2/»yP^y r^ sin (|) cos >}» + /72 X^w V cos^ <}.. 

The expression for dMt reads 

dM[ = y-6e ('^x;,^,^^ - a„^U_,U'y ~ a„ U'y) r dr. 

We Ccin represent this in the form 

- fl» (Ul + 2t/yA'(/y + A'^p] f dr 
or 

where 

A'rfA/t = - ^6oa„ (?A't/^^y + 2t/yA't/y + A'f/prrfr. (2.83) 

Substituting here the necessary e25)ressions from eqs.(2.77) and (2.82) /81 
and integrating, we obtain 



2k BR 



00 
or, setting mt = mt + A 'nit, we obtain for A'mt the e35)ression 

We now determine the expressions for PxXi and py\i with the aid of 
eqs.(2.73) and (2.75), assuming X to be small in ccaiparison with |j, (which holds 
for jx s 0.15): 



(2.85) 



82 






3 Su L j-L 2- 



X — ' ~^^ J^ !^ 



2 Q3l/?4 H 



(2.86) 



Obviously, the additional terms in the expressions for the flapping coeffi- 
cients and in the expression for the torque do not depend on X . 

Equations (2.86) show that the smaller the value of (j,, the greater will be 
the influence of distortion in the induced velocity distribution. This becomes 
understandable if we recall that the smaller the translational velocity, the 
greater should be the induced velocity caused by an aerodynamic force of the same 
magnitude. The fact that the quantities Px^i and pyXi do not depend on X greatly 
facilitates the calculation of the flapping coefficients and forces from the 
variable induced velocity. The calculation in this case is performed in the 
same manner as for a constant induced velocity, except that ao , a^ , and bi are 
calculated from eqs.(2.78), (2.79), and (2.80) and the expression for mt from 
eq.(2.48); to the obtained value, we add the term A'm^ calculated from eq.(2.85). 

ANALYSIS OF OBTADJED RESULTS 

11. Blade Flapping 

To illustrate the point, we made various calculations of a rotor in curvi- 
linear motion. 

Design data: 

rotor diameter, D = 14 m; 

loading coefficient, a = 0.065; 

static moment of blade relative to axis of flapping hinge, S^.h = 

= 1^1-2 kg • m; 

angle of blade setting, cp = 3°, a^o = 5.?. 

All calculations were performed for the regime of autorotation. 

Figure 2.1? shows the variation in flapping coefficients, calculated for 
a constant value p. = 0.3 and different values of the ratio ^ (rotation in /82 
transverse direction). We see from Fig. 2. 17 that, on rotation to the left, 
{ — 5— > q) bi increases while ai decreases, i.e., the cone described by the 
blades is deflected to the right and forward. In rotation to the right 
f — s_ < o), we ha,ve a decrease in bi and an increase in a^, which indicates de- 
flection of the cone to the left and backward. The coning angle ao slightly in- 
creases with rotation to the left and decreases with rotation to the right. 

83 



Consideration of the change in the induced velocity distribution leads to 
an even greater change in bi at an angular velocity Q^. 

Ciorve bi calculated for a variable induced velocity is plotted in Fig. 2. 17 

as a broken line. For 
practical purposes, we can 



ao.i^i.bi.flmln 















- 


„ 


- 


S" 

■3% 














l< 


































/ 


/ 


















- 










^ 


/- 






^ 


S^ J 








flo 


i— 








"" 




/ 








=- 




^ 


■r-' 


u 














/ 














^ 


S 












/ 


























"-^ 






- 


// 


// 


/ 


b, 








— 


















/ 
































/ 




































/ 






Si 








































\ 


■< 




"^ 


— 














p>min 


(^ 


^ 


r 






L 
















> 




^/ 














"V 


■n 


\ 


N 


s 










^ 


/" 


^ 


^' 








-0. 


oiy 


^0.03, 


'^0.02 -0.01 







0.01 


0.02 


O.L 


7J 


0.0^ 


El 
((I 


/ 




y 








"■ 




40 














-- 






V 




/ 




/ 


/ 








1 

- -z' 

-.7" 










- 








/ 


'/ 












/ 


/ 














- 














/ 


















disregard the changes in ao 
and ai due to variable induced 
velocity . 

Figure 2.17 also gives 
the curve for the values 



Pml n ■ 

shows 

Pmln 



— vai 



^2 



which 



a marked decrease in 
for rotation to either 



Fig. 2. 17 Flapping Coefficients in Transverse 
Rotation of Craft (iJ. = 0.3). 



left or right. This means 
that the reserve of blade 
overhang decreases markedly 
toward the lower arresting 
device. The blades will pass 
lower, the greater the angular 
velocity of the roll and the 
smaller the value of y • 

Figure 2.18 gives- the 
same curves for the case of 
rotation of the craft in 
longitudinal direction. In 
conformity with the foregoing, 
diardng rotation in a dive, ai 
increases and bi decreases, 
i.e., the plane of the blade 
tips is displaced backward 



and to the left, whereas during rotation in pitch it is shifted forward and to 
the right. Curve Pnin shows that nose-down rotation causes the greatest de- 
crease in p„i^, i.e., the blades pass very far below the plane of rotation, 
whereas pitching rotation in this sense is most favorable. 



M 



Figures 2.17 and 2. IS pertain to craft whose control is not acconplished by 
means of an automatic pitch control mechanism (for exanple, by ailerons and 
rudders, by deflecting the hub of autogiros with direct control). Both 
Figs. 2. 17 and 2. IS indicate that, for such craft in curvilinear motion (for 
exanple, during a sharp bank or going into a dive), the change in P„i„ may be 
much greater than its variation over the entire speed range of the craft in 
rectilinear flight. This should be taken into consideration in selecting the 
position of the lower arresting device of blade flapping. On a helicopter, the 
pilot, in deflecting the automatic pitch control, reduces the deflection of the 
blade cone in curvilinear flight; thus, the reserves of the angle toward the 
lower arresting device do not diminish so severely. 



84 



Oo,a,,b,,li„i„ 



■^ 
X 


< 
.b, 




— 












7* 




















^ 


























,y 


^ 












6" 
5' 


^ 




- 


"o 






y^ 






<-• 




N 










\^ 




, 


^ 






^ 








y 


y 






\ 


^ 




- 




/ 










3% 
2° 




< 






A 












- 






\ 


N 


/ 




'Pmin 






/ 
















■-•- 


^ 






















































■^ 




N 


>v 




















X 


























7 


/ 










■^ 


;n^ 






















^ 


■^ 


—J 





0.05j-0.0t 


-0.03 


-0.02, 


'-0.01 







0.01 


0.02 


0.03 


ao^ 


■\ 














/ 




1 
-1° 






























/ 


/ 
































r>o 
























- 




</ 


-' 






- 


























tO 






















/ 




































/ 














1 






















/ 


f 


































_ 








1 
-J" 






1 















783 



«2 



Fig. 2. IB Flapping Coefficients in Longitudinal 
Rotation of the Craft ((jl = 0.3). 




"fai 



:i. I 1 I I iTi 



Fig. 2 .19 Change in Angle of Attack and Thrust Coefficient 
of a Rotor, as a Function of the Angular Velocity of 
Rotation of the Craft at Constant iJ. = 0.3 • 



m 



85 



12. Effect of Curvilinear Motion at Autorotation 
of the Rotor 

In the aerodynamic, calculation of a rotor in autorotating regime, eq.(2.48) 
is used for determining X, -which permits finding the value of X, if mt = is 
assimed and cp, iJ., CI,., and Cly are known. However, in practice this quadratic 
equation becomes very cumbersome after substitution of the values of ao, ai, and 
bi. For determining the value of X it is more convenient, after plotting the 
dependence of m^ on X, to read from the graph the value of X at which mt van- 
ishes. When constructing the plot of mt, the coefficients ao, a.x, and b^ should 

be determined from eqs.(2.37), (2.38), 

and (2.39). 

K-gure 2.19 gives the values of /84 
the angle of attack of the rotor and the 
thrust coefficient t, obtained for a 
constant value y. = 0.3 at different 
angular velocities of rotation of the 
craft . 

Figure 2.19 indicates that, to 
maintain a constant value of |j, at con- 
stant flying speed in the presence of 
angular velocity, the magnitude of the 
angle of attack must be changed and the 
more so, the larger the angular velocity. 
When the craft rotates to the left, a 
constant value of ij, can be maintained 
only by increasing the angle of attack 
over that in rectilinear motion, whereas 
in rotation to the right the angle of 
attack must be reduced. This is ex- 
plained physically by the fact that, in 
rotation to the left, the rotor disk is 
Inclined forward (in addition to being 
inclined to the right), which reduces the angle between velocity and plane of 
rotation of the blade tips in space. In rotation to the right, the opposite 
occurs. It can be asstimed that, to maintain a constant value of [j., the angle 
between the velocity of flow and the plane of rotation of the blade tips must 
remain constant at any angular velocity Q. 

Figure 2.20 gives the values of the angles of attack as a function of |j,, 
obtained at different angular velocities of rotation of the entire craft. It 
can be seen that, if rotation of the axis begins and proceeds at a constant /85 
angle of attack (i.e., angle between velocity and a plane perpendicular to the 
rotor axis), then steady autorotation will occur only at a new value of p,, dif- 
fering from the former. Thus, in rotation of the craft to the left at a constant 
angle of attack iJ. will increase and in rotation to the right, it will decrease. 

If in addition to the angle of attack also the flying speed is kept con^ 
stant, the rpm of the rotor will decrease on rotation to the left and will in- 
crease on rotation to the right. 



a" 












- 




\ 


\ 


S 


^ 




- 




















\ 






fS 








L 












\ 














\ 
















V 










\ 




\ 






10 








\ 




\ 






\ 


\ 




\ 














V 




\ 










\ 


\ 












\ 




\ 




^ = + 0.05 


s 










\, 




\ 
























\ 




\ 




Vl 


>, 
















\ 






N 








^ 
















\ 








N 


. — 


& 


=n 


^ 


^ 
























\ 








""V 


->« 












*" 


"i 


















N 


11 








0.2 








C3n 


. 






0.1 




^ 


A 


















52, 


■ — n/iK 


^ 






























(ti 


I — 1 
















— 


— 



Fig. 2. 20 Angle of Attack of Rotor 
in Autorotation. 



86 



To illustrate this phenomenon. Fig. 2. 21 gives the revolutions of a rotor at 
angles of attack of a = +4.5°, 2.8°, and 1.5° ((io = 0.25, O.3O, 0.35 in recti- 
linear motion) as a function of the velocity of rotation in transverse direction. 

The number of revolutions in rectilinear flight is taken as no = 200 rpm. 

The rpm in the presence of rotation 
nQ is obtained from the following 




considerations : 



=9,55 



Ks=9.55 



Kcos g 
Vcosa 



-0.05 -COi^ -0.03 -OM 



Fig. 2.21 Rotor Revolutions in Auto- 
rotating Regime, as a Ftinction 

0. 



of 



(U 



ras = /to — ■ 

The value of i^q is determined 
from Fig. 2. 20 at the intersection of 
the ordinate 01 = const with the curve 
a = f(iJ.) for the corresponding angular 
velocity . 

As a consequence of the above 
effect of an increase in angle of 
attack necessary for maintaining 
steady autorotation at a given |j,, 
there exists the risk of the rotor 
losing autorotation at high flying 



speeds in the case of rotation of the craft to the left. 

If the helicopter begins to rotate sharply to the left at large \i. while its 
angle of attack remains constant, the rotor revolutions will decrease rapidly 
and |jb will increase further. This is aggravated by the fact that the forward 
inclination of the resultant of the rotor, generated when the craft rotates to 
the left, will per se (against the will of the pilot) create a diving moment 
which tends to decrease the angle of attack even more. The latter circimistance, 
increasing the abrtptness of drop in rpm as the craft rotates, involves an in- 
crease in angles of attack with respect to the blade sections, which causes flow 
separation and marked increase in drag. 

Figiores 2.22 and 2.23 give the variation in angles of attack, calculated 
for sections of r = 0.5 at p, = 0.3, for the case of stationary flight and for 
different directions of rotation of the craft. As seen here, the angles of 
attack with respect to the sections increase markedly on left-hand rotation and 
nose-down of the helicopter. This is due to the fact that, in the cases under 
consideration, an increase in angles of attack at angular settings of the blade 
of i|( = 270° and i|t = 180° .(where the angles of attack of the sections are already 
large) is required to balance the moment of the Coriolis forces. 

Taking the variable Induced velo'city into consideration will always in- 
crease the variation in angles of attack with respect to azimuth. 



87 



a 



^'r 








r~0.5; fi=0.3 








7 


r- 


\ 
















Rotation to left 















-ff=+m 








/ 




\ 








IS 




















/ 


^ 


/ 






\ 


























f 


i 








\ 










Bo 


tation to ri 


iht 


. 


-4 




/ 




\ 




\ 














^ = -0.05 






1 




y 






\ 




\ 




w 














1 


/ 






\ 




\ 














/ 




j . 






I- 


\ 






\ 




< Without rotation 
X - — 1 


/ 


J 


1 










\ 


\ 


\ 




\ 


\ 




9=0 




'~~- 


n 


/ 


j 






\ 








\ 












1 


/ 




1 






I 




- 


s 






\ 










/ 


/ 


/ 








\ 




\ 






\ 






J 


/ 




/ 






















\ 




N 


\ 







/ 




/ 










\ 














"s 


N 


\ 


I 


/ 




/ 












\ 


















A 


\ 






/ 














\ 








— — . 





— 








^>. 


^ 




















— 



100 



200 



300 360 (f' 



Rig. 2. 22 Angle of Attack of Blade 
Section vs. Azimuth. 



oi"r 




"TTl i 1 i 


1 1 




7\ 


— 


^ 


i 








« 


— 


Roto 


tion in diving — ' 












1 '^y- nm 














\ 












w 

1 _. I 


tation — 


— 


— 




L 




^ 


^ 


\ 


V 


- 




— 


Without ro 


















V 


W 


- 


LIJJ 

Pitching 


- 




- 


i 


y 




— 


N 


' 

N, 




^y ^n/T! 










[h 


^ 


\ 


>- 


^ 


■^ 


1 










\ 


\ — 




^- 










— 




// 






— 








f^ 


n 


^ 


5 




—2 


-^ 


- 


y 


/ 


P 


— 


— 








V 




















-N 


■v. 


























^ 






- 


— 


" 


— 


— 
















^ 


V 





















^ 




100 


/ 






2-00 




--- 




3UU 




3tL 








H 


\ 










\ 






















— 


z^ 










L 








_ 


1 





Pig. 2. 23 Angle of Attack of Blade 
Section vs. Azimuth. 



Based on vdnd-tunnel tests /86 
of blades with two profiles, of 
which profile Mo.l has higher /87 



values of c, 



and O'er than pro- 



file No. 2, we plotted the curves 
of the reserve of autorotation in 

Fig. 2. 24 (Ref.ll): 



a — Qq— tan" -^=/(a — Oo), 

Cy 



where a^ is the angle of attack 
at which Cy = for the profile. 
The diagram shows that, in the 
case cp = 3°, the range of angles 
of attack, at which an accelerat- 
ing moment is produced on the 
blade element extends from 0?^ = 
= 4° to ffr = 15° for the profile 
No. 2, and is much greater for the 
profile No.l, reaching a value 
of 30°. 

In profile No. 2, during left- 
hand rotation of the craft with 



a ratio of 



Q, 



CJU 



= 0.05 at n = 



= 0.3, a decelerating moment will 
he produced even on the section 
r = 0.5, over an appreciable por- 
tion of the disk. To maintain 
autorotation at given |j,, an even 
larger angle of attack of the 
rotor is required; if the angle 
of attack is kept unchanged, the 
rK>tor rpm will drop and the flap- 
ping motion will increase greatly. 

This brings us to the conr- 
clusion that a constant (accoiKi- 
ing to Fig. 2. 21) decrease in 
jTotor rpm on left-hand rotation 
of the helicopter and an increase 
in rpm on right-hand rotation 
will take place only up to some 



small value of 



U) 



whose magni-- 



tude is determined exclusively 
by the aerodynamic blade charac- 
teristics. At large values of 



88 



A-dg- tan-' ^ 




o. 



U) 



Fig. 2. 24 Curves of the Autorotation Reserve of Blades, 
the rotor rpm on rotation to the left will drop more abniptly - even as far 



as loss of autorotation - as a result of flow separation, whereas in rotation to 
the right the rpm vri.ll first cease to increase and then, at large values of 



U) 



-, begin to decrease. 



Thus, for a rotor with blade profiles of large autorotation reserves one 
can safely permit a much greater angular velocity than for a rotor with poor 
blade profiles. 



13- Behavior o f the Re sult ant of Aerodynamic Forces 
dji Cur vilinear HeUicopter Motion 



/88 



According to general considerations, the resultant in the case of curvi- 
linear helicopter motion is deflected in a manner similar to the deflection of 
the cone described by the blades in space. Thus, the resultant lags on the side 
opposite the rotation and, in addition, is deflected in a perpendicular direc- 



tion by an amount proportional to the ratio 






The lag of the resultant causes, relative to the center of gravity of the 
craft, a moment counteracting the rotation. This constitutes a damping moment 
which is larger, the greater the angiilar velocity of rotation of the craft. 

The inclination of the resultant in perpendiciilar direction in the case of 
lateral rotation causes a change in angle of attack, whereas in the case of 
longitudinal rotation the inclination of the craft is to the right or to the 
left. 

Figure 2.25 gives the variation in the quantity -7=- characterizing the 

angle of inclination of the resultant in the lateral plane, as a fimction of the 

89 



















s 








































1 

0.07 
0.06 
0.05 
0.01 
0.03 
0.02 
0.01 
















/ 


- 


- 
































/ 
































/ 
































/ 


































/ 
































/ 


































/ 




































/ 






















0.05 


-mv -0.03 -0.02 -0. 


^.OZ 
-0.03 
-0.0V 
-0.05 
-0.05 
-0.07 
-008 




0.01 


0.02 


0.03 


0.01 


& 


















































/ 




































/ 




































/ 




































/ 




































/ 




































V\ 



































Fig. 2. 25 lateral Inclination of 
the Resultant as a Function of the 
Velocity of Rotation in Transverse 
Direction. 



of the rotor with respect to the degree 
cients . 



angi^Lar velocity of banking — s_ (at 



CD 



p, = 0.3). 



The formula for the lateral force 
in rectilinear motion does not yield 
the values of S close to full-scale 
values (owing to the poor convergence 
of the calcxilated flapping coefficients 
to the real coefficients); however, the 
variation in lateral force as a func- 
tion of the angular velocity is cor- 
rectly given t^r eq.(2.62). 

The slope of the curve -^ = 



n. 



cu 



curve bi = fl 



■j is close to that of the 



This circumstance 



can be used for an approximate calcula- 
tion of the danping forces and moments 
of variation in the flapping coeffi- 



The magnitude of the conponent of the danping moment acting in the plane /89 
of angular velocity of the craft is readily determined from the formula 



M = Tby. 



where 



y = distance between center of hub and center of gravity of the craft; 
6 = angle of lag of the resultant in rotation. 

According to eqs.(2.80) and (2.86), we can take, for n s 0.2 



(0 \ 



Y 2 Qrt/?4(i ) 



B2(S2+-|-|.2) 



K,. 



Here, K^ is a coefficient taking into account the change in the flow coef- 
ficient \ during rotation of the craft. The coefficient K^ can be determined 
from experiment. If no e35)erimental data are available, we can take K^ = 1. 

The intensity of interaction, expressed by the slope of the curve S/T as a 
function of ^^M, may decrease on further increase of this ratio above a certain 
value, owing to flow separation in the blade sections. 

As a result of the above-described phenomenon of decrease in rpm and hence 
in thrust during autorotation while the craft rotates to the left (with a rotoi- 
of right rotation), the damping and thus also the control 1 ability of the craft 
is less in the case of left-hand rotation than in the case of right-hand rotar- 



90 



tion. 

The use of an optimum supporting blade profile extends the permissible 
range of the angles of attack of the blade sections so much that, with proper 
arrangement of the helicopter, the critical magnitude of the angular velocity of 
rotation of the helicopter will not be reached in actual service. 



EFFECT OF HDTOR PARAMETERS AND HUB DESIGN ON FLAPPING AND 

DAMPING OF THE ROTOR 



Blade flapping and deflection of the resultant in ciorvilinear motion of a 
helicopter are affected by the characteristics of the rotor itself, which 
"Changes its aerodynamics under these conditions. 

Below, we will examine the effect on flapping and danping of the rotor, the 
moment characteristics of the blade profile, its transverse centering, and hub 
design. 

Let us take a hub with a flapping con^^ensator, with kinematic dependence of 
the angle of blade setting on the flapping angle - such that the angle of blade 
setting decreases with increasing flapping angle. 



14. Rotor with a Profile , Having a Variable Center of Pressure 

Until now, we discussed a rotor having blades with profiles of constant 
center of pressure and with a flexural axLs coinciding with the center of pres- 
sure. Let us now examine a rotor 
having blades with a variable center 
of presstire. 

Recalling the variations in the 
angle of attack distribution of the 
blade sections with respect to azi- 
muth as they occur in curvilinear 
motion, it is easy to show that, if 
the coefficient of the moment rela^ 
tive to the flexural axis of the 
blade c„ depends on the angle of 
attack of the section, then the 
aerodynamic moment producing blade 
twist will vary in relation to its 
angular position. This, as a re- 
sult of blade twisting, will cause 
a change in the flapping motion and 
in the position of the resultant. 
Let us si:ppose that the flexural 
axis is located aft of the aero- /90 
dynamic center and that the profile, 
at aj. = 0, has a diving moment 




Fig. 2. 26 Diagram of the Effect of 
Coriolis Forces Producing Blade 
Twist in Curvilinear Motion. 



91 



(c„o < 0)» i«e«, vdth increasing a^ the center of pressure of the blade section 
shifts forward and the diving moment c„ decreases. 

As an exanple, let us examine the case of left-hand rotation of the craft. 

The character of the angle of attack variation of the blade sections, for 
rotation to the left, is plotted in Fig. 2. 22. For the blade in forward position 
{•if = tt), the angles of attack decrease almost to zero whereas for the blade 
located aft of (f = O) they increase appreciably. Therefore, the forward blade 
is tvidsted in diving, i.e., the angle of setting decreases, whereas the rearward 
blade is twisted very little (with the velocity U^ being identical in both posi- 
tions) . 

To balance the angles of attack of such blades, providing zero hinge moment, 
the cone of the blades and hence the resultant should be deflected to the left. 
This additional inclination of the resultant to the left, occiirring in left 
rotation, decreases the resistance of the rotor to rotation of the craft, i.e., 
danping, and may cause the helicopter to bank at large angular velocities. This 
effect produces pressiore on the control stick directed towa]7d the side of bank- 
ing. 

The above discussion shows that, if the flexural axis of the blade is lo- 
de 
cated in front of the aerodynamic center, i.e., if " < 0, the danping moment 

of the rotor resisting rotation of the craft, increases. 

15. Effect of Blade Centering 

If we assume that the center of gravity of the blade section is located at 
some distance bf from the focus of the profile (positive bf - backward), then 
during rotary motion of the rotor axis a couple, produced by the Coriolis and 
aerodynamic forces, arises on the blades. This is shown schematically in 
Fig. 2. 26. 

The ejqpression of the moment producing the blade twist (a positive moment 
will twist the blade toward an increase in angle) will have the form 


Substituting dKQ and dKQ by their expression from eq.(2.7), we obtain /91 



M^={2b,m (2y(i) sin 1^ -|- 2^(o cos tj)) r dr. 



It is obvious here that the moment varies periodically. 

The angle of twist of the blade, \mder assunption of constant mass, a value 



92 



of t), , and torsional rigidity, is determined ty the formiila 



C Mdr 2 bimRs /Q o > 



Assuming that a linearly twisted blade is equivalent to a blade with a 
constant angle of setting equal to the angle of setting of the first blade at 
the section r = 0.75 R, we find 

bcmR^ 






Let us derive an additional periodic variation in angle of setting 



where 



Av= Vj cos (j) -|- V, sin tj>, 



Qv - 



Vi=>^— ; v, = yi-^, >\=0,61 



6j m/?3io2 



The periodic variation in angle of setting leads to a variation in inclina- 
tion of the cone during rotation. 

The flapping coefficients take the form (for constant induced velocity) 

1 






-\-A 



2 '^ 



^ ^ ' B'ilm+ — (i.2 j 



-/I 



B2 + — p.2 



(2.87) 



If the center of gravity is located aft of the aerodynamic center (bf > 0, 
A > O), an additional inclination of the cone and hence of the resultant to the 
side of rotation of the craft will occur. If the e.g. is ahead of the a.c, 
then the additional inclination increases the danping moment of the rotation. 

The angles of twist Vj, and v^ are easily determined if the dynamic twist of 
the blade in rectilinear flight is calciilated and the angle of twist Vq is /92 

93 



knovm (Ref.6). The relation between these angles is determined by the formulas 

2 Qy 

(2.88) 



V,=Vn 



v,=v„ 



2 Q^ 



16. Rotor with Flapping Compensator 

Let us now examine a rotor whose blades change in pitch cp, as a function of 
the flapping angle p. There are many methods of acconplishing such kinematics. 

Figure 2.27 shows one of the methods 
of changing pitch in relation to the 
flapping angle p (turned flapping 
hinge), where cp varies in accordance 
— . with the law: 




Axis of blade 



Axis of flapping 
hinge 



Fig. 2. 27 Diagram of Blade with 
Turned Hinge. 



to the case ij, > 0. 



The flapping angle P, in this 
case, is the angle between the blade 
axis and the plane of rotation. Mot 
wanting to conplicate the results, we 
will take the case n = 0. The varia^ 
tion in flapping motion obtained for 
regimes with p. = can be extrapolated 



Let the craft be inclined in space at a constant angular velocity having 
the conponents Q^ ^'^'^ ^y* 

The velocity conponents of the blade element will take the form 



U^=\wR- 



■ r -6- + 2„r cos i]j — 2 ^r sin ip. 



Assiming cp = 9o - P tan a^ , we construct, as above, the ecjuation of moments 
relative to the horizontal (flapping) hinge for B = 1: 



Y 2y 



+5i+K'->-H(f+^)+ 



Y ^x 



-j- -^ -^ COS t|j sinil) — 2 — sin--j< — 2 — cos^. 

4 <0 4 0> * CO u 



(2.89) 



The particular solution of this equation has the usual form 

p = flo — 'J'l cos <|> — &, sin (p, 



94 



while the coefficients of the series have the values 

1 



'^=v(i+4-) 



1 -|- tonoj 

4 

2^ 



I u V Y / ' <o \ Y 7jl+t«2ai 

L u) \ Y / 0. V Y VJl +t«A2oi 



131 



(2.90) 



Equations (2.90) show that, in the case of a turned hinge, the deflection 
of the cone described by the blades and hence the resultant will take place at 

a smaller phase shift than in the case of a 
conventional hub. The absence of phase shift 
TABIE 2.3 means that, in transverse rotation, there should 

be no change in longitudinal inclination of the 



Y 


8 


6 


4.6 


0], deg 


45 


37 


30 



resioltant, i.e., at 



Q 



cu 



ai = and, conversely, at 



i^ = 0, we should have 



ou 



0, bj 



0. 



This condition is satisfied if 



tanai=— r 



(2.91) 



The values of a^ , shown in Table 2.3, are derived with respect to the 
value of Y • 



TABLE 2.4 





o, = 


V 


= 8 


Y = 4.6 






ai = 45° 


o, = 


oi = 30° 


«! 


_S3, 




Qy 


Qy 

—1.74 — 


Qy 

- 1.74 — 

(0 


*I 


Qy 







Qy 





1 *' 
tan-^ 

«1 


45° 







30° 






It is of interest that the condition of absence of phase shift tan a^ = 
= y/8 yields the same magnitude of damping (lag of the resultant) as for a rotor 



95 



vd-thout flapping conpensator. For exanple, for longitudinal rotation we ottain 
the values indicated in Table 2.4» 

It is ob-vious that the absence of phase shift in transverse rotation at 
constant angle of attack ensiires maintenance of jo, in the aut ©rotation regime, 
i.e., prevents a decrease in rotor rpm or in controllability during rotation. 

B3T0R FLAPPING IN CURVILINEAR MOTION OF THE ROTOR AXIS 
AT VARIABLE ANGULAR VEIOCITY 

Below, we will derive formulas for determining the flapping coefficients 
of blades in the presence of uniformly accelerated and harmonic oscillation of 
the rotor axis. For simplicity, we will take the case where |j, = 0. It is shown 
that, for both laws of variation in angular velocity of the rotor axis, the /94 
flapping coefficients and hence the longitudinal and lateral forces of the rotor 
will vary by the same amoimt - proportional to the angiilar velocity - as in the 
case of uniform rotation of the rotor axis. Furthermore, terms appear that de- 
pend on the angular acceleration of the rotor axis. 

17. Uniformly Accelerated Rotation of the Rotor Axis 

Let us first examine the case of rotation of the rotor axis in pitching at 
variable angular velocity'"" 

^> = ^^- (2.92) 

Let us substitute t = -|- and put k = -^. We can then write 

2y = % (2.93) 

We then derive the eijqjression for forces and their moments relative to the 
flapping hinge. 

The velocities in the blade section, assuming v = const. |J. = 0, Hy = ktjf , 
Cl^ = 0, and B = 1, can be obtained from eqs.(2.23T and (2.24): 



Uy=^y.i^R-i-k<]ir COS <l/~r -^ . 



(2.94) 



Then, the thrust moment takes the fonn 

Af^=-i-j^»Qa„(cpt7J + ^^t/y)rc/r= (2.95) 



^ By tiorning the coordinate axes through an angle \|fo , all results obtained below 
can be extrapolated to the case of rotation in any direction. 

96 



2 \ 3 ' 4 4 rfij; ' 4(0 ^ J 



The moment due to the Coriolis force is 

M^,=---2/^^^k'^sln^. (2.96) 

The inertia force of rotational acceleration reads 

dQy 

ajs=m r cos t|» dr, 

where '^^ 

dQy d{ki/) 

= •=«U). 

dt dt 

The moment of this force is 

R 

M]^= l"/?iAo)COs<})r2rfr=/^^^;feiocos'j;. (2.97) 

u 

The equations of moments relative to the flapping hinge, after canceling 
by 1^,^ • (ju2, will take the form 

Suu (2.9s) 

The particular solution to this equation has the form 

{)= flj, — a^'if cos •;) — i^i'^ sin ip — ^i cos tjj — afj sin ijj. (2-99 ) 

Let us find the derivative of p with respect to \|/ : /95 

£L= ajijj sin 4' — a, cos (j> — 6,i]> cos ')' — &, sin ij;-!- c, sin 'J' — rfj cos t)!; 
.£?L= flj,]) cos <})+ 2ai sin ([» -f frjij) sin i|j — 2&, cos <J<4- <^i cos '^ + ^1 sin iji. 

dtj;2 

We next substitute these values into the left-hand side of eq.(2.98), 
yielding 

-^a,<j>sln<]) —biifC0Si(-{-(2ai ^^iH — ^CJ^sintj) — 

-(2*,+^H,+-Jrf,)cos.l> + «o = Y(|-+^)+ (2.100) 

+ — cos <!> — 2 — <}» sin<j) + -^ — >}. cos <1< — -'"•^ 

<d CO 4 w /». 



'v.v,"'' 



97 



Since eq.(2-100) is an identity, then, hj equating the coefficients of like 
terms, we ottain the following system of equations for detemiining the coeffi- 
cients of the series 



4 ' 4 a)' 






(2.101) 



(2.102) 



Prom eqs.(2.10l) we obtain 



a, = — - 



*,: 



c,= 



y (1) 



(0 V ^2^' 



, 12 /fe 
rfl = . 



(2.103) 
(2.104) 
(2.105) 
(2.106) 



Thus, the solution for P can be written in the following form: 









1 



64\ 

— ) COS (^ - 



y 0) 



sintj). 



(2.107) 



It is easy to see that the quantity — i- is none other than the ratio of 

the instantaneous angular velocity n. = kijf to the velocity od. Thus, the coeffi- 
cients of the first terms in eq.(2.107) are analogous to those previously ob- 
tained for Hy = const, namely 



«!= , Oi = — 

Y" 






The terms containing sin \lr in the expression for P are derived from the /96 
influence of inertia forces generated as a result of nominifomi rotation. In 
backwajTd rotation, the inertia forces tend to lift the blade which is in the rear 
position; this causes a change in the flapping motion and a decrease in angles 
of attack so as to attain equilibrium. In so doing the ajd-s of the cone tilts 



98 



to the left. 

So far we have investigated the particular solution for eq.(2.98)» charac- 
terizing forced oscillations of the blade. Let us now examine the general solu- 
tion of eq.(2.98) without the right-hand member, i.e., the equation 

Setting y = 8 in the particular case, we find the solution in the form 

p=C,e-*-fC2<l.e-*. (2.108) 

The general solution of eq.(2.98) then becomes 

X 



p = C,e-^ + C,<^e-^ + Y (-y + ^) + 



8ki> kii \2k C 2.109 J 

-| ~ COS (jj -f- — !- sirn|j sin ijj. 

yiii (I) Y** 

The values of the coefficients C^ and Cg are fo\ind from the initial condi- 
tions ^i = 0; P = 0; p' = 0: 






(2.110) 



As we see from eq.( 2.108), the terms containing Ci and Cg decay extremely 
rapidly; thus, in one revolution (t = 2Tr) the degree of perturbation diminishes 

tenfold: e" * = e'^'" « 0.002, te"* « 0.012. 

This furnishes a justification to use only the particular solution of 
eq.(2.107)» neglecting free oscillations of the blade, a procedure also confirmed 
by experiment. 

A conparison with experiment showed that, xinder static operating conditions, 
the induced velocity distribution over the disk has a substantial influence on 
flapping; the refined formiilas for the flapping coefficients are given elsewhere 
(Ref.8). 

For a rotor with a flapping conpensator, the flapping motion of the rotor 
is determined from the formiolas: 



/ X . So \ 1 SwM 1 . 



1 + -J-Wo, ^"^ 1 + -— tono, 
4 4 



99 



^'^-TO-T'^^'OHi^' (2.113) 

c,^-[a,i,...,+±)-T,^l-±t^o,)+J-±,^o,]^^; (2.114) 

^.= -[a.(l-i-t.o.) + ^(w.+J-)+^]_J__. (2.115) 

A conparison of eqs.(2.1l2) and (2.113) for % and bi followed by conparing 
them with the previously obtained expressions (2.103) and (2.104) for a rotor 
without flapping conpensator, will show that the lag of the cone and hence of /97 
the resultant in a direction opposite to the rotation will be practically the 
same, whereas the inclination in a perpendicular direction will decrease. 

18. Harmonic Oscillation of the Rotor Axis 

Let us now examine the case in which the rotor axis executes a harmonic 
oscillation in space at angular velocity 

where p = -^ with v being the vibration frequency of the craft. 

Since danping of the free oscillations of rotary-wing craft is small, the 
harmonic law describes oscillations of the craft close to the true oscillations. 
We again obtain the expressions for the mcments of forces relative to the axis 
of the flapping hinge. The velocities in the blade section, assuming as usual 
M. = 0, B = 1, are equal to 

rfp (2.116) 



(J = Xisifi -\- Asin pi!^r cos iJj — r 
^ dt 



The thrust moment is 



u 

= ^bQa^R^.^(^-[-^-^^+ ^sinp^.cos4> \ (2.117) 

2 \3'4 4d(^' 4« / 



The moment of the Coriolis force reads 



My^= —2 f /ni4sln/7iiu)rsin(jj£/r= — 2/ , (B/\sin/7tjjsirnj;. (2.118) 





The moment of the inertia force of rotational acceleration is 
100 



^ d2v 



form 



Mjt^= \ in -^^rcos'<^rdr=I^^^ A pm cos p^ cos <^. 



The equation of moments, after canceling ty I^.h • o)^, is vo?itten in the 
Tr!' + 4'7r + f^= sinp^sinii.+^— sin^cos<l) + 



i_A^cos/76cos<b + Y('— +-^V (2.119) 

(0 \ 3 4 / 

The solution to this equation (assuming that we can neglect the free motion 
of the blade, according to the foregoing) is foimd in the form of 

B^flo — <^iC0s(/7ilj— -({j) — felsin (pi|j — il») — al cos (ytTtji + ilj) — 6 1 sin (piji + <])). 

Substituting this solution into the equation of motion of the blade, we 
find the values of the coefficients: 



Id 



,. A f(/''-2/» + f (P-I )--^^(P-I) 

Ol = — 

to Y'^ 



798 






' to V^ 

-rr (P + 1)2 + (P2 + 2p)2 
lb 

,. yl 8 4 4 2 

-^(/'+l)» + (/'2 + 2p)2 



Disregarding powers of p greater than the first (since p does not exceed 
0.03 - 0.04) and expressing sin (pilf ± if) and cos (pijr ± li ) in terms of the product 
of the form sin pi|r cos ■i/ , cos pil( sin \|; , cos p^lf • cos t , sin piji • sin i/ , we 
obtain 

p==a.4-i- Asln;;tl.cost5.-f— +1 ) — pcos p^-cos'b-^ 
Y « \ Y •' " 

A 4 A 

-| sin/^ijisint}!-! /? cos p'ji sin ij). 

to Y ^ 

101 



Recalling that 

pA cos /?tp 



i4sin7^|»=2y, 

dQy ^ 1 dQy 

dif <a dt 



we find that the flapping motion, in the case of harmonic oscillations of the 
rotor axis in space, can be represented in the form 



^=".+[f^-(^+') 



dQy I 



dt 0)2 



COS li-f- 



A^_^±^n (2.120) 

[ u ' Y tii "'^ J 

Thus, the longitudinal inclination of the cone of the rotor during rotation 
in a longitudinal plane vdll be 

a.= -^^+(^ + 0^-^. (2.121) 

and the lateral inclination becomes 

'" 0) Y d^ "^^ (2.122) 

These expressions, relative to the magnitude of the terms proportional to 
the angular velocity of the rotor shaft, coincide with those previously obtained 
for uniform and uniformly accelerated rotation and thus can be used, in the 
general case of helicopter motion, for determining forces and moments of the 
rotor, danping the helicopter motion. 

For a rotor with a flapping con^Densator, eqs. (2.121), (2.122) vary propor- 
tional to eqs. (2. 102) - (2.106). 

CHARACTERISTICS OF ROTOR AERODYNAMICS DETERMINED /22 

BI HINGED BLADE ATTACHMENT 

Hinged blade attachment has a substantial effect on the aerodynamics of the 
rotor; therefore, an -understanding of the role and pl::ysical meaning of flapping 
motion will help the reader toward a better study of the characteristics of 
rotor aerodynamics. These questions are presented below. Fxirthermore, a siirple 
graphic derivation of formulas for calculating the flapping coefficients is 
given. 



102 



19. Physical Meaning of Blade Flapping 

The equation of tlade flapping in steady rectilinear flight can be repre- 
sented as 

where 

?r^; m..= -^rrfr. 



"oo ~ 2/h.v, ' ^"-^ J d? 

Here, S,,. n is the mass static moment of the blade relative to the flapping 
hinge. 

As shown above [see eq.(2.33)]» the e^qjression for — ■::- contains the flap- 

dr 

ping angle p and the angular velocity 7^ ; this demonstrates that the flapping 
motion relative to the flapping hinge is danped by aerodynamic forces. 

Owing to appreciable aerodynamic danping [with linearization of the equa- 
tion, i.e., on the assimption that Cy = acoaj., the average (per revolution) coef- 
ficient of J? in eq.(2.34) is equal to —i— B^ « 1 - l.?], the natural oscilla^ 

tions of the blade rapidly die out and the flapping motion of the blade becomes 
a forced oscillation due to the thrust moment. Since the natural frequency of 
the blade is close to the rpm of the rotor [the average (per revolution) coeffi- 
cient of p in eq.(2.34) is equal to 1], the blade reaches its maximum oscilla- 
tion anplitude ipon a variation in thrust moment with the frequency of the rotor 
revolutions, i.e., with respect to the first harmonic. 

The correlation between the anplitudes of the second harmonic of flapping 
and the thrust moment is by approximately a factor of 10 less than for the first 
harmonic. Therefore, despite the fact that the anplitude of the second harmonic 
of the thrust moment is high, blade flapping occiu?s mainly with respect to the 
first harmonic. 

Thus, the bulk of the flapping motion of the blade is described by the 
equation 

EJ = o.(, — a,cos.]j— ftisinij). (2.124) 

Let us substitute eq.(2.l24) into the equation of flapping motion (2.123). 
The left-hand side of the equation is equal to ao: 

^+P=ao. (2.125) 

The equation of flapping takes the form 

103 



m^^^^a^ 






(2.126) 



(p=270'' 



It follows from eq.(2.l25) that, in each section of the blade, the svrai of 
the inertia force of flapping and of the conponent of the centrifugal force /IQQ 
normal to the blade axis is proportional to ao and is a constant, remaining un- 
changed ipon rotation of the blade al- 
though the flapping angle of the blade 
changes. This means that the first 
harmonic of the moment of inertia forces 
relative to the horizontal hinge is 
equal to zero. Therefore, as shown by 
eq.(2.l26), the thrust moment of the 
blade relative to the flapping hinge 
should be the same at all azimuths. 
Herein lies the basic characteristic of 
a rotor with flapping hinges and the 
physical meaning of blade flapping with 
respect to the first harmonic: The 
blade moves about the horizontal hinge 
so that, as a resiolt of the redistribu- 
tion of aerodynamic forces over the 
blade caused by the flapping, the thrust 
moment relative to the horizontal hinge 
does not change at all azimuths. 




Fig. 2. 28 Distribution of Thrust 
over the Blade Radius at Identical 
Magnitude of Thrust Moment Rela- 
tive to the Horizontal Hinge. 



20. Redistribution of Aerodynajiiic Forces over the 
Rotor Disk due to Flapping 

Equality of the magnitude of the thrxist moment of the blade relative to the 
flapping hinge at every azimuth will not resiolt in blade thrust, calculated only 
with consideration of the first harmonics of flapping which are the same at all 
azimuths, since the distribution of thrust over the radius changes from azimuth 
to azimuth (Fig. 2. 28). However, owing to flapping of the rotor with hinged 
blade retention, the first harmonic of the change of blade thrust decreases 
steeply. 

The blade thrust depends on the flapping motion mainly with respect to the 
additional relative flow normal to the blade axis"" produced during flapping of 
the blade elements, which changes the true angle of attack of the element. The 
changes which introduce first-harmonic flapping into the distribution of true 
angles of attack over the rotor disk are appreciable. For exanple, the addi- 
tional vertical velocity of the air AUy = a^r of a blade element at azimuth 
i|r = 90° and of the same element at azimuth ;j; = 270° is the same in magnitude but 
opposite in direction. However, owing to the difference in the horizontal comr- 
ponents of the relative flow, the true angle of attack of the element decreases 



■5^ For sinplicity, we will call the velocity of the air normal to the blade axis 
the "vertical" velocity. 



104 



little at \|f = 90° and increases much more at i|t = 270° . This explains the local 
increase of the true angles of attack of the blade sections in the region t = 
= 270° and the occurrence of flow separation at high flying speeds for a rotor 
■with hinged blades (Fig. 2. 29). 

Above, we determined the relative vertical velocity of the flow at azimuths 
i|r = 90° and i|; = 270°. This was found equal to, respectively, -aircu and aircu. 
The ej^jressions have a sinple explanation. 

Figure 2.30 shows a rotor whose blades have different flapping angles at 
azimuths i|r = 0° and f = 180°, i.e., the axis of the cone of the blades is de- 
flected backward (ai ^ O). Here, the blades have a maximum vertical velocity 
with respect to absolute magnitude on passing through azimuths ijf = 90° and /lOl 

i|( = 270° since the blade, in the same time interval At = ^ , is vertically 



cu 



displaced by the largest magnitude (p > n) . At azimuths i|r 
the vertical velocity of the blades is equal to zero. 



0° and i; = 180°, 




^'^/SO' 




f =90° ■,270° 



Fig. 2. 29 Variation in Angle of Attack Fig. 2. 30 Displacement of Blade Section 
of the Blade Section with Respect to Relative to Plane of Rotation on 
Azimuth, due to Blade Flapping. Blade Turning. 

Thus, a change in vertical velocity and, consequently, in true angle of 
attack and blade thrust at azimuths ilf and \|; + rr will take place on variations in 

TT "^ 

the blade flapping angles at azimuths i|r + -^ and % + -^ and vice versa. Bear- 
ing this in mind, it is easy to understand how the rotor flapping will vary if, 
for some reason, a cyclic change of the true angles of attack takes place or an 
additional moment relative to the flapping hinges appears on the blades. 

For exanple, if because of blade twisting or for some other reason the 
angles of attack of the sections increase to a maximum at azimuth \|f and decrease 
maximally at azimuth % + tt, then an additional flapping motion of the blades is 



105 



established so that the blades occupy the lowest position at azimuth i|( 



TT 

2 



and, when flapping upward, reduce the true angles of attack to a value at which 
the condition of constancy of thrust moment relative to the horizontal hinge is 
observed at all azimuths. The highest position of the blades is at azimuth ^ + 

+ -O- after which they drop, restoring the diminished angles of attack. Along 

with the variation in flapping with respect to the first harmonic, the forces H 
and S also vary (Fig.2.3l): 

A// = rA3sin6; 

~^S=T^'pcos'!^. 



It was shown above that, despite the large first harmonic at velocity U, 
the first harmonic of the variation in blade thrust with respect to azimuth is /103 
relatively small, since it substantially decreases because of the flapping. The 
second harmonic of blade thrust is larger and the third smaller than the first 
harmonic . 




Plane of 
rotat ion 



Fig. 2. 31 Variation in Flapping and 
Longitudinal and Lateral Forces due 
to Dynamic Twist of the Blade with 
Respect to the RLrst Harmonic. 



P 
P" 



m 



h.h 



0.10 



-0.10 













/ 


/ 




^ 


_ 










- 








/ 














'"h.h 


^ 


k 








-- 


> 


N 




\ 


s. 








> 


v^ 


-- 






\ 






\ 


\/ 
















/ 


1 


\ 




^ 






/0(N 


\ 


\ 




200 










— 


_- 




- 










\P" 










\ 


\ 













7102 



'300 ^°^ 



Fig. 2. 32 Variation in Flapping Angle, 
Angular Acceleration of Flapping, and 

Thrust Moment of Blade Relative to 
Flapping Hinge as a Function of Azimuth. 



The second harmonic of blade thrust causes second-harmonic flapping motion 
of the blade 



Afi-= — fljCOs 2()j — ^2 sin St]), 
which is equalized by the moment of inertia forces 



I^^ »^ (^^ + Ap) = /,., ">= (3^2 cos ^ + 36, sin 2^) 



(2.127) 



(2.128) 



106 








\ 


\ 














\ 












^ 




- 








/ 










/ 








/ 


r 








■^ 


-^ 






[> 


- 






^ 


^ 


^ 


— 


^'• 










s 1 


N 












\ 


^ 




























20(S 


\ 






300 




r 
































/ 




\ 










/ 










\ 






/ 








\ 


^ 


^ 







Fig. 2. 33 Variation in Angular 
Velocity of Flapping, Angle of 
Attack of Section at r = 0.975, 
and Blade Thrust as a Function 
of Azimuth. 



and creates some redistribution of aero- 
dynamic forces of the blade with respect 
to azimuth, which is less extensive than 
for the first harmonic. 

The higher harmonics of flapping 
are very small and have practically no 
effect on the blade aerodynamics. 

The graphs in Figs. 2.32 and 2-33 
are an illustration of oior statements 
on blade flapping and variation in aero- 
dynamic forces with respect to azimuth. 
The diagrams were obtained by rough 
calculation, on the assutiption of uni- 
form induced velocity distribution over 
the rotor disk and without consideration 
of elastic oscillations of the blade 
which affect the magnitude of the upper 
harmonics of flapping and blade thrust. 

The calculation was performed for 
the following initial data: 



K=0.30; t^^QAQ; a=-9.4°; Mo = 0.6; 



= 0.9; ;fe=0, /, 



k.v," 



= 0. 



The curves show the kinematic characteristics of flapping P,'— rr-, ^—, 



W 



dr 



thrust, and thrust moment of blade t^, m,,.^ , and angle of attack of section cfp 
at r = 0.975. 

We see from this exanple and from Table 2.11 that, beginning with the 
second harmonic, the flapping coefficients markedly decrease and, beginning with 
the third harmonic, the decay coefficients of blade thrust diminish. Thus, the 
angle and angular velocity of flapping as well as the angle of attack of the 
blade section vary mainly with respect to the first haraionic, i.e., with the 
frequency of the rotor revolutions. The second harmonic becomes manifest in 
angular acceleration of the blade, whereas the blade thrust and its moment rela- 
tive to the flapping hinge vary mainly with respect to the second harmonic . 

21. Approxim ate Derivation of Formulas for Flapping Coefficients 

On the basis of the properties of blade flapping, described in Subsection 19, 
we will derive approximate expressions for determining the flapping coefficients 
ai and bi obtained in Subsection 4« For sinplicity, we wiU take B = 1 and /1G4 
will disregard small terms of the order of y, so as to obtain expressions with 
an accuracy to |j, . 

On the basis of the constancy of the thrust moment at all azimuths, we will 
equate the thrust moments for azimuths differing by 180° . This method permits 



107 



a better definition of the mechanism of equalization of thrust moments by means 
of flapping, under different conditions of blade flow at azimuths differing by 
180°. 

The angle of backward tilt of the axis of the rotor cone ai is determined 
from an examination of azimuths \lf = 90° and ) = 270°; the angle of sideward tilt 
of the axis of the cone toward the side of the advancing blade (\jf = 90°) is de- 
termined from azimuths ilf = 0° and 'ij = 1S0°. 

The superposition of the translational velocity of flight on the rotary 
motion of the rotor is responsible for the different operating conditions of the 
blades at azimuths 90° and 270°. At azimuth i|i = 90° the velocities are added 
and at azimuth ijr = 270°, subtracted. Therefore, the coefficient a^ is equal to 
zero during static operation of the rotor and increases with an increase in fly- 
ing speed V (or ti = ^ ^°^ "^ 

At azimuth ^ = 90°, the relative flow in the plane of rotation is equal to 
Ux = (jaR(r + IJ.). Here, in the region of large velocities, the backward displace- 
ment of the axis of the rotor cone causes a lifting of the blade and a decrease 
in the vertical conponent of the relative flow Uy = u)R(X - air), which reduces 
the true angles of attack of the sections. 

At azimuth t = 270°, the relative flow in the plane of rotation is small, 
while the vertical velocity and the true angles of attack of the sections in- 
crease: Ux = cuR(r - |i); Uy = u)R(X + air). 

Let us then construct the equations for the elementary thrust moment, take 
the integral from r = to r = 1 at both azimuths and, equating the results, 
find the expression for ai. We can also equate to zero the moment of the thrust 
difference at azimuth \|i = 90° and t = 180°: 



where 



Hence, 





a, = 2ix(Acp-^x). (2.130) 



Owing to the velocity difference of the oncoming flow at azimuths j = 90° 
and \|r = 270°, the quantity ai will vary even at the same change in angle of at- 
tack or vertical velocity for the blade at these azimuths. For exanple, vipon an 
increase in angle of attack of the rotor, equal vertical velocities appear at 
the blade sections at azimuths 'if = 90° and t = 270°. To have the blade thrust 
moment increments at these azimuths identical, the angles of attack of the 

108 



sections at azimuth \|; = 90° should he decreased further and, at azimuth i|; = 210° , 
increased again. Ob-viously, this viill occur xxpon an increase in a^. 

This is an important property of a rotor mth hinged blades: Upon an in- 
crease in angle of attack of the helicopter owing to an increase in a^, the 
longitudinal force H increases and a destabilizing moment appears relative to /105 
the center of gravity of the helicopter, causing an even greater increase in 
angle of attack; the helicopter is statically iinstable with respect to the angle 

of attack. 



AUy^-Vd^ 




Fig. 2. 34 Variation in Velocity 

Uy as a Function of the Coning 

Angle ao • 



We should note that a^ does not depend 
on the inertia characteristics of the 
blades, since a^ equalizes the aerodynamic 
"asymmetry" in rotor operation. 

The presence of the coning angle is 
responsible for the difference in vertical 
velocity Uy of the air relative to the 
blade, at azimuths t = 0° and t = 180° 
(Fig. 2. 34). 



For a blade in the forward position 
(t = 180°), the velocity of the air is directed from the bottom ipward; to reduce 
the true angle of attack the blade, on passing the azimuths 90 - 270°, is lifted 
tpward. Diur-ing the second half of the revolution the blade drops, which inr- 
creases the true angles of attack. Thus, the axis of the rotor cone is dis- 
placed laterally, toward the side of the advancing blade (| = 90° )» 



Let us now derive the expression for the coefficient bi 
ponents of flow around the blade sections are equal to: 



The velocity com- 



at azimuth i|/ = 0, 

at azimuth il; = 180°, 

U^^wRr; L/y = u)R {l — bir-\-\).ag). 
Equating the thrust moments of the blade at these azimuths, we obtain 



*i = ^!^ao- 



(2.131) 



The coefficient bj equalizes the aerodynamic "asymmetry" caused by the 
presence of ag. Since ao depends on the mass characteristic of the blade Yj i't 
follows that also bi depends on y . 

22. Eff ect of Monuniformity of the Induced Velocity 
Field on the Flapping Motion 

Next, we will define the variation in the flapping coefficients ai and b^ 

109 



for the case in which an additional vertical velocity appears on the sections 
and an additional periodic moment relative to the flapping hinge acts on the 
blade : 

Af/y=— t/iCOs.]j — f/jSint)'; (2.132) 

AM=— Micostj) — TWaSintp. , -_„n 

The blade thrust moment is the only moment able to balance the additional 
first-harmonic moment caused by a variation in the vertical velocity and in the 
moment AM. 

The linear thrust of the blade receives an increment owing to a change in 
the flapping coefficients by a quantity Aa^ and Abi. In this case, the equation 
of flapping has the form 

R 

^drr = /^^^^i>?ao~M^cos'!^-M^sin'!^. (2.134) 

In conformity with eq.( 2.134) » we can examine the following equalities: /106 
l\dTr] +^2=(Urr| -M,; (2.135) 

lUTr] +M,=r{dTr] -M, (2.136) 

or, in dimensionless form, 

\o /4.=90° \o Ji 

d\^4-7d7\ +--i-^/w,=-LfN7rfA -^S-^. (2.138) 

\0 /^=Q° \0 /(ti=180» 

The equality (2.137) can also be described differently: 

Ceo J l-U'-A=90'' Vrf'-A-270"J Y/v..W'"2 (2.139) 

or, expressing in the form of 
dr 

110 



"1 ^- M.\ 

j Y/w.t,'"^ (2.137) 

' 41=270' 



dt /dt\ ( dt\ , ( dt\ . , 

ilCf)/''"^ = -^^- (2.141) 

The physical meaning of eq.( 2.141) is obvious: The flapping hinge moment, 
varying with respect to the first harmonic, is equalized by the moment of the 
first harmonic of thrust'"". 

Henceforth we will use the equality (2.139) and, accordingly, the 
equality (2.1^2): 

Lf[(^) -(^-L) l7rff=-2— ^vW.. (2.1^2) 



"oo 



and determine only the flapping coefficient increments. /107 

Let us examine the azimuths i]; = 90° and i|f = 270° (K.g.2.35): 



-^Af^) =_(Aa/-fi7,)(r + p.); 

-^a(^) -=(Aa,F+t7,)(7-,.); 

a^ \dr /^=270° 
J-fAf^) ~^(^) 



= — 2Aa,r2 — 2f/2r. 



We 



will give the expressions for [ ^'^ ■] and ( — :^) (with an accuracy to p.^); 

\ dv \ ^ dr /i 

1 /dt\ - , _ 

— ( -^ ) =r2 ai — |xX — 2r|X9; 
a<» \dr /\ 

aXdr/i ^ o^ L W/a.v ~ \^A=oJ "" ^''"^^ """ ^'^^ ~ 
— [ipr2+ r (X +bir — i>-ao)]=riiao — r2bi. 



However (-^^) is not equal to (-^r) - (-^) „ since 
\ dr /i ^ dr /av ^ dr /|=90 



the term 



-|j.air in the expression ( — — -j is the coefficient of sin^'i]; and does not per- 

V dr ^|=go° 

tain to the first harmonic. 



Ill 



ujS(r*u) 



u)R(aa,r*U^) 



ujUfr-fi) 




ujR(Aa,ri-U^) 



Fig. 2. 35 For Deriving the 
Expressions for Aai. 



1 

'J 



From the equality (2.139) we obtain 



Aa, = — 4 U.r'^dr 



.P 



Y/h.W<-2 



Af, 



(2.143) 



Next, let us examine the azimuths ijr = 
and ilf = 180° (Fig. 2.36): 



From the equality (2.142) we obtain 



A&i=4 U.r^dr '- ..,, 

Using eqs.(2.l43) and (2.144), we then derive the formulas for determining 
the flapping coefficient increments, -with consideration of a nonuniform in- /108 
duced velocity distribution over the rotor disk. 



Vcosoc 




u)R(&b,r-u,) 



Fig. 2. 36 For Deriving the EjqDressions for Abi , 



In first approximation, the induced velocity distribution can be described 
by the equation (see Fig. 2. 6) 



v{r, ']()=i) -far CDS']*. 



(2.145) 



Since the positive direction of the additional vertical velocity AUy is 
from the bottom \ip and that of the induced velocity from the top down, a comr- 



112 



b, 



/ 




16 




tg 




-=r^^ 






"^^ 


/ 




^y^'''^ 


J 




/ 




""'"^ 


T'^ofi 











parison of eqs.(2.1A-5) and (2.132) will 
yield 



,=0. 1 






{2.im 



0.1 



Substituting eq.( 2.146) into eqs.(2.l43) 
and (2.144), we obtain the sought expres- 
sions: 



Fig. 2. 37 Variation in b^ as a 
Function of \i,. 



Aai = 0; 



A^i=4a j rMr = a. 



{2.1hl) 



Thus, the backward deflection of the axis of the rotor cone vd.ll not change, 
whereas the lateral deflection will increase by an angle numerically equal to 
the increment of the relative induced velocity at the blade tip in both forward 
and rear positions. If we assxmie that a = v"", i.e., that the induced velocity 
at the leading end of the disk is equal to zero and at the trailing end equal to 
double the mean value, then 



Aftj=T) = 



ia 



4fl2 K(Jt2+ \2 ' 

while the total value of backward deflection of the cone axis will be 



(2.148) 



*i = TrH'«o + - 



<0 



4B2/( 



^2+X2 



(2.149) 



A longitudinal tilt of the induced Arelocity field also affects the magnitude 
of the longitudinal and lateral forces of the rotor. Let us derive the formulas 
for detennining h and s: 






(2.150) 

/109 

(2.151) 



Equations (2.149) - (2.151) can be used at p, > 0.1 - O.O5. Therefore, in 
Fig. 2. 37 which gives the curve of bi as a fimction of |j,, the sector from p, = 
to iJ, =0.1 contains a broken curve, laid approximately through the points p, = 0, 



"- The quantity a, as related to the flight regime, can be determined from data 
given elsewhere (Ref .25). 



113 



at which "b^ = and ij, = 0.1. 

By means of eqs.(2.143) and (2.144) we can also obtain the approximate 
expressions for determining the flapping coefficient increments during curvi- 
linear motion of a heliccpter, which were derived in Subsection 4* 

METHOD OF CALCULATING THE AERDDYWAMIC CHARACTERISTICS 
OF A ROTOR FOR AZMUTHAL VARIATION OF BLADE PITCH 

23* Equivalent Rotor Theory 

It will be shown below that a rotor whose blade pitch changes cyclically 
with respect to the first harmonic 

? = ?o — ViCOSip — 9iSina), (2.152) 

can be regarded in the aerodynamic design as a rotor with a constant pitch equal 
to cpo, but with a different angle of attack. On this basis, the method of de- 
termining the aerodynamic characteristics of a rotor with a pitch variable in 
azimuth is called the equivalent rotor theoiy. 

The equivalent rotor theory furnishes an explanation for the mode of varia- 
tion in rotor characteristics with deflection of_the automatic pitch control 
mechanism. The formulas for calculating cpi and cpx in relation to the angle of 
deflection of the automatic pitch control and the kinematic characteristics of 
the rotor hub are given in Subsections 25 - 28. Data published earlier (Ref .15, 
14) were used in presenting the material. 

First, let us examine the problem formally: 

Substituting eq.( 2.152) into the equation of flapping and, for sinplicity, 
retaining only the first harmonics, a series of transformations will yield 



Below, in Subsections 23 and 24, only the pitch conponents cpo , cp 1 , and 91 
will be contained in the formulas so that, for sinplicity, we will omit the sub- 
script "0" of cpo • 

From eq.( 2.153), the following formulas are obtained for the flapping /HO 
coefficients : 



114 



ao=Y[Y+T^(^ + l*'>-irH= (2.154) 

«i=^ — 7 — (>'+Y'p-^i"^)-^«; (2.155) 



i-T<^^ 



3(1 + -,2) 



(2.156) 



It is obvious that, on making the substitution X^q = \ - ^itx, eqs.(2.154) 
to (2.156) can be rewritten in the fonii 



^»=^[^+T^(^+'''^]= (2.157) 



' 2 ■ 

*i = — : -, r '^oP' + ?i 



(2.158) 



(l+'Y'^') (2.159) 



A comparison of these formulas with eqs.(2.40) for a rotor with constant 
pitch readily shows that ao and the first terms of the expressions for ai and bi 
coincide, provided that both rotors have equal |j. and cp, and that the X of the 
rotor with constant pitch is equal to X^q. • Henceforth we will denote all quanti- 
ties pertaining to a rotor with constant pitch by the subscript "eq" or "e" 
( for e quivalent ) . 

The coincidence of the formulas enables us to detemiine the flapping coef- 
ficients of a rotor with variable pitch from the_foniiulas for the flapping coef- 
ficients of a rotor with constant pitch, adding cpi and cpi : 

^o = ^Oe; (2. 160) 

«i = «ie-'Pi; (2.161) 

ftj = 6,^+9i. (2.I62) 

In so doing it is necessary to satisfy the conditions of equivalence of 
the rotor with variable pitch and the rotor with constant pitch: 

X — ^,[i = X,; (2.164) 

9=?e- (2. 165) 



115 



Now we are convinced that the following relations are satisfied: 

_ _ _ (2.166) 

(2.167) 
or 



Acpi/^ = (9iCOS-|--|-», sin.j»)f/^=Af/y. 



(2.168) 



Actually, on the basis of eqs.(2.l60) - (2.165), we represent both sides /111 
of the equality (2.168) in e^^^anded form: 

— A'f t/x= (?i cos (jj + (p, sin ij)) (7+ [J. sin •]>); 
At/y = AX — r (Afli sin 4) — A*, cos ij>) — (xACo cos t}) -f [xAa, cos2 <J) + 
-f l^Afti sin <^ cos ijj=yj[j„— r ( — 9, sin ')' — ?, cos <)>)— 

— cpiti. cos2 (j) -f (p,(A sinip cos ijj == (^1 cos ip +"^1 sinip) (r + [X sin <);). 

It is obvious that the equality (2. 168) is valid here. 
It follows directly from eq.(2.l68) that 

Uy Uy^ ^(f Uy 

°r=? + -^ = ?e + A?-h— + — '=cp,+-^ = a,^. (2.169) 

Thus, the angles of attack at all blade sections for the rotor with variable 
pitch and for the rotor with constant pitch equivalent to it are equal. 

likewise, we can show that 

^^(^\ . (2.170) 

dr \d? J^ ' 

t, = t,^; (2.171) 

y_. * (2.172) 

t — tj. 

Equations (2.168) - (2.170) show that a decrease or increase in linear 
thrust, produced by a change in pitch of the blade at a given azimuth, is due to 
a decrease or increase in Uy at the same azimuth when calculating on the basis 
of the equivalent rotor theory. 

At equal thrust coefficients, the relative induced velocities are also equal 



ta t.Q 



116 



-V,, (2.173) 



from which it follows, based on eqs.(2.l63) and (2.164), that 

X as (AO — v, 

X — X,= ti,(a — aj = 9i(j,; (2.174) 

«e = « — ?i- 

We represent the e^^jression for — ^ in the form [see eq.(3.56)3 

dr 

1^— ir|+'-,„0;. (2.175) 

Using eqs.(2.l68) and (2.170), we find /112 

dr \ dr I U, + "^^".J^^ \ dr lu7~U?k ~ 

~ (-^\ (y, COS 6 + ^1 Sin d>); ( 2 . 176 ) 

?<i' = 9+j — ^^((picostp + yisin 6); 
'"*-^^J'^^J'~3'^^^ = '"*c--^j(?'^°^'!'+^i^'"^)'=''I'J(:f-)'^'^^- (2.177) 



Since, for a rotor with flapping hinges, the value of the integral 

r f B rjrdr is constant at all azimuths, the integral with respect to i|r must be 

equal to zero. Consequently, the average per-revolution magnitude of the torque 
coefficients of the rotors is identical: 

'"t='«t,. (2.178) 

dq 
However, at equal azimuths the values of ,_ , qx, and mt^ for both rotors 

are not the same and the rotors have a different variable conponent of flapping 
motion about the drag hinge. 

Let us now derive formulas for deteraiining the coefficients h and s of the 
rotor with variable pitch from the corresponding coefficients of the rotor with 
constant pitch: h.^^ and s,, . On the basis of eqs. (2.161), (2.I62), (2.171), and 
(2.177) we then obtain 

AA^= — ^^ /\pcos<li-j-A^4.sin(l)= — ^^ X 
X[(<PiC0si}i — (p,sin4i)cos i})-|-(^iCOsi}i+^iSiin)))sin ({)]= —^4, ^j; (2.179) 



^e' 



117 



As^= — ti/ Apsinii — A^i^ cos '^ = t^ <d^. 



(2.180) 



Consequently, 



^h 



2r. 



-^Z'iv 



S = Se + ^e?l- 



(2.181) 
(2.182) 



Finally, let us define the relation of the coefficients of forces in the 
velocity axes. It follows from eqs.(2.172), (2.174), and (2.181) that 



ty^f. 



^^ «i /a + A = /,a^H- /?^ + ^, Aa + A A = ^.,^+ ^, 9, - t^^^ = t^^ 



(2.183) 
(2.184) 



Equations (2.183) and (2.184) show that a rotor with different cpi and cpi at 
identical p,, X^^, cp has identical ty and t^ . Consequently, at equal p,, ty, t^ 
the rotor with a cycHc variation of pitch and the rotor with a constant pitch 
have equal cp, X^jj, oi^^f but different a. This characteristic of a hinged rotor 
manifests itself in that, at equal ty and t^ (at equal f Hying weight, speed, /113 
and altitude) but at different ^i (different centering or angles of stabilizer 
setting), the helicopter will have different angles of attack and angles of 
pitch. This is shown in Fig. 2. 38: The rotor with constant pitch and the rotor 
with variable pitch, at equal p,, ty, tx have equal ag(j but different a; conse- 
quently, the helicopter with a deflected automatic pitch control mechanism in 
the same flying regime will occupy a new position in space. 



t,S 



Plane of rotation of rotor 
with pitch variable 
in azimuth 




Plane of rotation of rotor with constant 
pitch and plane of equivalent 
rotor for 



aj^(x,=ar^-<p,= a. 



Fig. 2. 38 Angles of Pitch of Helicopter at Same Flying Regime 
but Different Deflections of Automatic Pitch Control. 

An TTppm-tant consequence of eqs.(2.l78), (2.183), and (2.184) is the possi- 
bility of mathematically determining the interdependence of the coefficients |j.. 



118 



ty, tx, nit irrespective of whether or not the_rotor has a cyclic variation of 
pxtch with respect to azimuth since, for any cpi and cpi, the coefficients ty, t^, 
nit do ^'^^ change. This property of the rotor greatly sitiplifies the aerodynamic 
design of a helicopter. 

The above-derived formulas of the equivalent rotor theory will remain valid 

even if they are not derived from eq. (2-153) 
and even in the absence of assi:mptions of uni- 
formity of the induced velocity field (without 
discarding higher harmonics of flapping) and of 
other assunptions. Consequently, also here 
transformations based on the equivalent rotor 
theoiy will hold. The higher harmonics of flap- 
ping and the loads acting on the blade in the 
thrust plane, for a rotor with pitch varying 
as a function of the first harmonic and for a 
rotor with constant pitch, are identical if the 
conditions of equivalence of the regimes 
(2.163) - (2.165) are satisfied. 




Fig '2.39 Displacement As of 
the Flapping Hinges Relative 
to the Plane of the Equiva- 
lent Rotor. 



The equivalent rotor theory is not appli- 
cable in the case of widely spaced flapping 
hinges, since relative to the new reference plane, i.e., relative to the plane 
of the equivalent rotor, the blades execute an additional displacement As 
(Fig. 2. 39) together with the flapping hinges, which does not occur when calcu- 
lating a rotor relative to the plane of rotation and is not taken into account 
in design formulas . 



ip=const 




h\ 







If ''Const 




^ . cX=oli(at ipfO)Of 



a -(X, y [fi^o 





Fig. 2. 40 Reconstruction of the Aerodynamic Characteristics 
for a Rotor with Pitch Varying in Azimuth. 

Finally, for a rotor with constant pitch all dimensionless characteristics 
are defined xsgon prescribing three quantities (jj., \, cp or any other three quanti- 
ties), whereas for the rotor with variable pitch five quantities (iJ., \, cp, /114 
^i> 9i °r 3-^ other five) must be known for determining the dimensionless 



119 



characteristics in the related axes. 

Thus, it has been proved that the calculation of a rotor mth v-ariable 
pitch can be replaced by the sinpler calculation of a rotor with variable pitch, 
i.e., stipulating equivalence of the fljdlng regimes ( 2.163) - (2.165), viith sub- 
sequent conversion by the above formulas. 

The sequence of calculation is as follows: 

From the quantities ii, A., cp, 91, cpi which are known for the rotor with 
variable pitch, we find p,e»^o>9e* 



We then determine a 



o.> 



■'I.* 



mt,. 






From the conversion formulas, we find ao, aj,. 



The equivalent rotor theory is often used in determining the aerodynamic 
characteristics of a rotor from graphs. If the graphs are constructed for a 
rotor with constant pitch, their change for a rotor with variable pitch will be 
as shown in Fig.2.40. In the graphs for the angle of attack at 9 = const (upper 
plots) the curves of t are equidistantly shifted by Aof = cpi, and each point of 
the curves of h is shifted by ha = cpx 'to the right and by Ah = -tcpi downward. 
On the graphs for rotor pitch, at_Q' = const (lower plots) the marking of the 
angles of attack is changed (for cpx f^ 0» each curve corresponds to an angle of 
attack greater by cpi), and the curves of h, in addition, are shifted by Ah = 
= -tcpi. The graphs of mt, ty, t^, ao, and of higher harmonics of flapping a^, 
b„ (n = 2, 3» •••) are modified like the graphs of t, whereas the graphs of s, 
ai, bi are modified like the graphs of h. On the graphs of the aerodynamic 
characteristics in the velocity axes (the plot on the right in FLg.2.40) for a 
rotor with cpx ^ 0, the marking of the angles of attack is also changed. 



Plane of 
equivalent 




7^ 

Plane of rotation 



Fig. 2. 41 For Determining the Position of the 
Equivalent Rotor Plane. 



Let us now derive formulas correlating the characteristics of the rotor 
with variable pitch and its equivalent rotor with constant pitch, on the basis 
of geometric relations. 

Figure 2.41 gives a side view of the rotor and two blade sections at azi- 



120 



muths 90 and 270° . If we draw a plane turned about the blade axis through an 
angle Z^i|f_9oo = -cpi to the plane of rotation, then the blade pitch relative to 
the turned plane will be identical and equal to the mean value of pitch per 
revolution. This plane is the plane of the equivalent rotor. The angle of /115 
attack of the equivalent rotor a, = a - ^)1. If cp^ ji^ 0, an analogous picture is 
obtained on viewing the rotor from the azimuth i|f =0, i.e., the plane of the 

equivalent rotor is turned, relative to 
-r;- the plane of rotation, through an angle 

9i in the side plane of the helicopter. 



oTT 



Plane of equivalent 




^ Plane of rotation 



Fig. 2. 42 Velocity and Elementary- 
Force Conponents of Blade Sections 
in Different Reference Planes. 



Thus it is obvious that, for a rotor 
with pitch varying cyclically with re- 
spect to the first harmonic, we can se- 
lect another plane of reference relative 
to which the rotor pitch does not change . 
Therefore, relative to the new reference 
plane we can detemiine forces, moments, 
and flapping of the rotor by formulas 
derived for the rotor with constant 
pitch. In so doing, it must be taken 
into account that the new reference 
plane has a different angle of attack 
and that the results of the calculation 
pertain to axes related with it and 
should be converted to axes related with 
the plane of rotation of the rotor. 



This constitutes the geometric meaning of the foraiulas derived above. 



The position of the aerodynamic force of the rotor relative to the velocity 
vector of flight does not depend on the selection of the reference plane; there- 
fore, its conponents on the velocity axes, i.e., lift and propulsive forces, are 
equal [see eqs.(2.183) and (2.184)1. 




fi 7 ~^=^ 



Fig. 2. 43 For Determining the Difference Uy - Uy 
at Azimuths i; = 0° and ijr = 180° . 



Let us now outline the changes occurring when calculating the elementary 
forces of the blade section on change-over to the new reference plane. 
Figure 2.42 shows the blade section at azimuth i|r . The section has a setting 



121 



angle cp\|i relative to the plane of rotation and a setting angle 9 relative to /II6 
the plahe of the equivalent rotor. The angle of attack of the blade section, 
i.e., the angle between the chord of the blade and the vector of the total rela- 
tive velocity of flow around the section U, does not depend on selection of the 
reference plane [see eq. (2.169)]. The relations between the conponents of U at 
small values of kp are equal [see eqs.(2.l66) and (2.168)] rU^^ - U^; Uy^ == Uy + 
+ Ux/icp or AUy - -/kpUx . 

As indicated above, the last expression shows that any decrease or increase 
in load per lonit length of the section due to a change in blade pitch at a given 
azimuth for an equivalent rotor with constant pitch is the result of a decrease 
or increase in Uy at the same azimuth. 

Let us define the reason for the variation in Uy at the characteristic azi- 
muths ilf = and 90°. At azimuth i|f = 0, Uy and Uy are equal: 



^.r'K-h)-.-'i^) 



¥e see from Fig. 2. 43 that, at \|f = 0, the value of p changes on changing to 
another reference plane by the same q-uantity as a so that a - P = a^ - 3e • This 

r, 77 - r dp /' dp A " 
means Uy - Uy^ = - r ^-^ - l^-gf-; J • 

If /\q) = J^^ cos t, then the plane of .the equivalent_rotor is inclined 
laterally relative to the plane of rotation by an angle cpi , on account of which 



Thus, when the pitch of the blade at azimuth ^ = changes by -"cpi^ a change 
to the equivalent plane in the calculation will lead to a decrease in Uy owing 
to a decrease in the flapping rate relative to the plane of the equivalent rotor 
by a quantity equal to cpir. 

If /kp = -9i sin i|f, then -||- = (-||-) and AUy = /kp = 0. 

At azimuth i/ = 90°, Uy is equal to 

If /^ = _9i cos iji, then at ^; = 90° -^ = (-g|-) , a = Qf^ , and AUy = Zkp = 0. 



122 



If Zkp = -cpi sin i|f, then a - o-, = 91 , -^ - \-^) = -cpi" and /117 

iJjy = \>.'fi + r-f, = 9i (f -f p.)= — A?t=90-Z7^^_9o. . 

Consequently, when the blade pitch decreases at azimuth cp = 90°, the change- 
over to the plane of the equivalent rotor produces the same decrease in aero- 
dynamic force as a result of the fact that Uy < Uy owing to a decrease in angle 
of attack of the equivalent rotor and an increase in flapping rate relative to 
the equivalent rotor. 

In conformity with Fig. 2.42, the formulas for converting the load per unit 
length in the blade section will be 



cir 



dt _/ dt \ 
d'r ~\ d'r )^ ' 



Thus, all fonnulas of the equivalent rotor theory are in essence only 
formulas for converting from one system of axes to another. 

24. D erivation of Formulas for a R otor with Flapping; Hinges 
as for a Rot or without Hing^. Conditions of 
Equivalence of ffi-n g:ed_and Rigid Rotors 

In the Glauert-Lock theory, when deriving formulas for the coefficients of 
forces, torque, and flapping, the flapping angle of the direction of forces, in 
space is reckoned from a plane relative to which the setting angle of the blade 
in rotation remains constant. Obviously the plane of the equivalent rotor meets 
these requirements. 

In this Subsection, we will derive formulas for the coefficients of forces 
and torque of a rotor, except that we conceive the hinged rotor as rigid relative 
to the axis of the cone described by the blades. In so doing, we will take the 
plane of the blade tips as the reference plane rather than the plane of the 
equivalent rotor. Relative to this new plane, the blade setting angle changes 
in rotation but there is no flapping; this sinplifies the expression for the ve- 
locity conponent of the flow past the blade Uy normal to the reference plane. 
Since Uy enters the expressions for elementary forces more conplexly than the 
setting angle, the formulas for the coefficients of forces and torque in the tip 
plane are sinplified. 

This method gives individual form-ulas applicable to the calculation of a 



-■^ Figxire 2.43 shows the displacements of the blade As relative to the plane of 
rotation and plane of equivalent rotor during a half -revolution of the blade; it 



is obvious that 



f' dpj, ^ dp 



\ 



dt y, dt * 

123 



rotor both with constant blade setting and mth a setting angle variable relative 
to the plane of rotation. 

Occasionally, approximate- expressions for the longitudinal and lateral 
forces of the rotor enter the aerodynamic calculations and especially the sta- 
bility calculations: 



h=ta^; 



tai; 1 



(2.185) 



Obviously, these expressions are valid if the forces directed parallel /118 

to the plane of the blade tips are 
equal to zero, i.e., if the re- 
sultant of all aerodynamic forces 
is perpendicular to the plane of 
the tips. 



bloil^ 




Plane of 



'^"^cosj^"'^ 



rotation y^ // 



VsLn(a.*a,) 



The obtained expressions for 
the coefficients of forces parallel 
to the plane of the blade tips are 
additions for refining eqs.(2.185). 



Fig. 2. 44 Velocity Conponents of Flow 
Past the Blade. 



Finally, we will derive various 
formulas while retaining the as- 
sunptions of the Glauert-Lock theory. 
Blade flapping can be taken into 
account only with an accuracy to 
the first harmonic. For a rotor 
with infinitely heavy blades (ao = 

= t-^ = g = 0, the coefficients of higher hannonics of flapping are also equal 

to zero) such formulas were derived by Lock (Ref.37)« 

Let us now derive these formulas . 

The velocity coirponents of flow past the blade in a plane normal to its 
axis (Fig. 2. 44, plane N) are the conponent parallel to the plane of the blade 
tips 



Ux, . ss (or -|- V cos (a -f ai) sin 6 ^ uj/? (r -j- (i sin <^) = wRUj 



"(K) 



(K)' 



and the component normal to Uj 



(k) 



i/„ 



'(K) 



\/ sin {a -f Oi) — 1) — V COS (a -f o-i) cos •]> sin ag : 



where 



= toR (X-f jifli — flojji, cos (]j)=co/?(X^|.j — flgf). cos ■\i) = wRUy 



^K)=>-+fliiA; 



(2.186) 



(2.187) 



(2.188) 



\f characterizes the velocity of the airflow through the plane of the blade 
tips. 



124 




Let us first examine a rotor 
having a pitch constant with re- 
spect to azimuth. In the section 
normal to the blade ajcis at azi- 
muth t, the angle between the 
plane of rotation and the plane 
of the blade tips is equal to 



P 



TT 



= aj. sin ^ - bx cos i|i 



Fig. 2.45 For Deteraiining the Blade Set- 
ting Angle Relative to the Plane of the 
Blade Tips. 



(Fig. 2.45). Therefore, the rela- 
tion between the quantities per- 
taining to the plane of rotation 
and the plane of the tips, in con- 
formity with Fig. 2.46, is as 
follows : 

?(K) = tp — P ^ =<p — a, sin <{) + &, cos <|i; (2.I89) 



^-^,K,^^A 



^y,K>=*^y + ^-P . =t/, + d/^ (a, sin <}.-*, cos <1>); 



4, +. 



Tw^T; 



Q(K)=Q — T'P „=Q — r(aiSinij.-6,cos<}.). 



(2.190) 

/119 
(2.191) 

(2.192) 
(2.193) 



The expression for the angle of attack of the blade section obviously 
should not depend on the selection of the reference plane. We will demonstrate 
this, after substituting eqs. (2.189) - (2.191) into the expression for the angle 
of attack of the section: 



'-'X 



Uy <fUx + Uy 



'fU. + Uy^^-U^^ 



+ + 



Ux 



Uj. 



u^ 



(,-....).,.,..,., ^_^^„^^_^^„_ 



(K) 



f/^ 



'(K) 



U^ 



'•(K)- 



(K) 



(2.194) 



The expressions for the coefficients of aerodynamic forces of a blade ele- 
ment in the plane N have the usual form (see Fig. 2. 46): 

("S"), =(SCos*(K) + c^pSin<I?(K,)Z7^«aoo(9(K)t7i + t7„ U^); 



125 



Ondtting intermediate conputations, we can give the final formulas: 

- flopr cos <!> i- flot^^ sin 2<1j i- i^^-^, ,, cos Z^l; ( 2 . 19 5 ) 

+ (2|iC^p^^r — aoo'f (K)^(K)(i) sinijj + Co (2!i.X(K)ao + 9(K)aoH.r) cos ij> -f- 

+ -J- aooao9(K)[i^ sin 2<j; — J- ^^ (c^^ + aoocg) cos 2t, ( 2 . I96 ) 



where c^p is the average value of c^p over the disk. 

To determine -~- we must use eq.(2.193): /120 

Jl.=(Jl) +(-^) (a, sin .J,- ft, cos i>). (2.197) 

dr \ dr Jw \dr /(k) ^^ ^ ' ^ 

Substituting cp, with respect to eq.(2.189) into the e^qjression for 
(-^i^l and integrating, we obtain 

Since, in reality, the rotor blades have flapping hinges, the condition of 
flapping motion 

I 

Y ndrj (2.199) 

is satisfied, which yields 



2m- 
' 3 



(V) +T^)=— V- ('+T'^)= ^2.201) 



1 + -^.^ 1 2 



^=-7^^Vv• (2.202) 



i(i + |.^) 



Let us derive the expressions for the coefficients of longitudinal and 
lateral forces parallel to the plane of the blade tips. 

126 



The elementary longitudinal and lateral forces are equal to 

(-^) =(—^\ sin<j» — ('-^^ On cos 0; 
V dr Ak) \ dr JiK) ^ \ dr /(«) 

(^) ^-(^) cos4.-(^) aosin^. 
\ dr /(K) \ dr /(k) \ dr /(k) 



(2.203) 
(2.204) 



Substituting eqs.(2.195) and (2.196) into these equations and integrating /I2l 
the elementary forces over the blade radius and azimuths, we obtain the follow- 
ing expressions for the coefficients of average per-revolution forces: 



Cxp.^l^ a„ 



S(K,= - — (ao [y W- -fr (1 +3|x=)] + 3X(K)(ao!^ — ^)} . 



(2.205) 
(2.206) 



With consideration of these expressions, the coefficients of longitudinal 

and lateral forces of the rotor 
are equal to 







(2.207) 
(2.208) 



Fig. 2. 46 Velocity and Elementary Force 
Conponents of the Blade Section for 
Different Reference Planes. 



Calculations show that use 
of the approximate formulas 
( 2.185) leads to an error in de- 
termining h, equal to 10 - 30^ 
(toward the side of a decrease, 
a larger figure always pertains 
to small thrust coefficients t) . 
The derivatives of the coeffi- 
cient h with respect to |jb, a, cp 
can be determined from eq.(2.185). 
Determination of the coeffi- 



cient s and its derivatives by means of eq.( 2.186) gives & resiilt differing 
greatly from calculations by means of eq.(2.208). Equation (2.208), just as 
eq.(2.6l) of the Glauert-Lock theory, at -uniform induced velocity distribution, 
only approximately determines s, but calculations by means of eq.(2.208) are 
closer to the experimental data than calculations by means of eq.(2.185). 

Practical calculations show that the value of the coefficient s in autoro- 
tation is close to zero and amounts to only a small portion of the value of the 
product tbi . At average values of m^ (horizontal flying regime) the value of s 
is smaller than that of tbi and we can roughly consider s = ^ tbi . At maximum 
power conditions, the value of s is equal to or higher than tb^ . 

The torque coefficient of the rotor is determined from the expression 



127 






(2.209) 



Since eq.(2.209) provides for integration -with respect to ilf vdthin limits 
from zero to 2tt, all harmonics of the expression for ( — —-) vanish on inte- 

gration. Therefore, it siiffices to substitute into eq. (2.209) only the center 
portion of i — —] : 

(K) 

•' .2 2 ,2 / 1 1 /, \;1 (2.210) 

Substituting eq. (2.210) into eq.(2.209) and integrating, we obtain 

We will demonstrate that mt = m^ . In fact, using eq.( 2.197), we find ra.^: 

viC ) 

... , /i22 




^ '"* (K) "'"^ J ^"' ^'" * ~ *' "^^^ *) ^'l' J "S' '^^'^' 





(2.212) 



It is known from the equation of blade flapping relative to the flapping 

hinge that the expression for the thrust moment of the blade f J —tzt rdr^ does 

not contain first harmonics ijr ; consequently, the integral with respect to ilf is 
equal to zero. Thus, the coefficient of the average per-revolution torque m^ is 
determined by eq.(2.2ll) (the instantaneous values of m^, are not mutually equal, 
i.e., m* 5^ mt ). 

It is also easy to prove the equality of the ejqDressions for mt [eqs.(2.2ll) 
and (2.48)] 'by taking their difference: 

128 



The expressions in the brackets are equal to zero, since they represent 

formulas [eqs.(2.201) and (2.202)] for determining the flapping coefficients aa 

and bi# Thus, it has been proved that m,. = m^ 

( k) 

The lift and propulsive forces of the rotor do not depend on the mode of 
calculating the conponents of. the resultant force of the rotor in the related 
axes, whether relative to the plane of rotation or relative to the plane of the 
blade tips. They are equal to (considering cos ai ~ 1, sin ai « ai; 

^y =/(K)Cos(a-^-ai) — A(„)Sin(a4-ai) = ^ cosa — Asin a = ^ • (2.213) 

(K) * 

^A-(K) = ^(K)Sin(a + a,) + A(K)COS(a + a,) = ^sina + Acosa = ^^. (2-214) 

The formulas derived in this Subsection are sinpler than those of the 
Glauert-Lock theory. They are also of interest in that they permit tracing the 
manner in which, and the factors by which, the formulas for calculating a rigid 
rotor (sinplest case) are transformed into foiroulas for a hinged rotor. The 
change in formulas takes place for the following reasons: 

1. Change in angle of attack of the rotor owing to deflection of the angle 
of rotation of a rigid rotor through an angle ai . In place of \ we introduce 
X, . = X + aiiJ, into the formulas for a rigid rotor: 



^(K) = aoi 



-+■ 



^(K) I V / i , 3 



(■H^f 



A(K)= — ^ H'^(K)?; 



S(K) = 0; 

2. Cyclic change of rotor pitch relative to the deflected plane of rota- /123 
tion of the rigid rotor: 

?(K)=<P-9,(K)C0S'1' -?i(K)Sin<l'. 

With consideration of the cyclic change of pitch, the fonnulas for a rigid 
rotor take the form 

'-='•- [V-¥+f(>+T''f 

«(K)= — - h'<)\?\>- ^/+'^ — ' 

S(K)=— — >-(K)?l(K); 

129 



m 



,,= -T^-(^-+T'^-N^)+^(i + ^^). 



where cpi = ^it ^>i - -^i ^o^ "the case in which the blade pitch does not 
change relative to the axis of rotation. 

We note that, in calculating a rigid rotor with variable pitch, it is im- 
possible to use a new reference plane relative to which the pitch is constant, 
i.e., the equivalent rotor theory. This is due to the fact that the blades, on 
rotating, do not lie in a new reference plane but actually leave it, i.e., per- 
form flapping motion relative to it, and the rigid rotor theory does not hold 
for the new reference plane. 

3. The coning angle of a rigid rotor in forward flight creates a cyclic 
change in the velocity con^jonent Uy of the section flow (see Fig. 2. 34) J AUy = 
= -|i.ao cos t. Furthermore, owing to the presence of the coning angle the blade 
thrust is projected onto the plane of the tips, stpplementing the forces h^^^ 
and S/^x . These factors further conplicate the formulas, so that they acquire a 
form which, after the substitutions cpi = -b^ and cpid^^ = ^i, coincides with 

that of eqs.(2.198), (2.205), (2.206), and (2.211): 



^(K) = 



"""[t 



) "fl(K)l^ 



H 



i+fi^^ 






2 ' 



fw--'-^(l+3(.^) 



'"t(K) = 



"" (1 A I 2 ?i(K)M'\ , aoii ,, 



x(aot^ + -|-?.,K,)]+^(l+l^^). 



(2.215) 

(2.216) 
(2.217) 



(2.218) 



4. Change-over from the plane of the blade tips to the plane of rotation /124 
according to the expressions 



h=--h^K)-'!-ia^; 



ntx 



tn 



t(K). 



(2.219) 



) 



130 



When using eqs.(2.2l5) - (2.218) for calculating a rotor with flapping 
hinges according to the condition of flapping motion [eq.(2.199)], ^o, ^ir^) > 
and 9l(l^^ will be equal to 



ao=Y 



\k) . '?l(K)f^ 



f^(l + !^=) 



'i'lCK)^- 



4go|j. 



2 '^ 



2f^ 



1 u2 

2 '^ 



x(^('<)-'-?i(K)l^+y ?)• 



X 



(2.220) 



Let us next examine a rotor with pitch varying in azimuth. 

Equations (2.215) - (2.218) without any changes are applicable also to 
calculating a rotor with variable pitch. 

In this case, the conditions of equality of the setting angles relative to 
the plane of the blade tips are 






(2.221) 



where 9l(^^ > 9i(ic) f ^^^ ^o* a^s before, are detemdned in conformity with the 
condition of zero moment at the flapping hinge in accordance with eqs.(2.220). 
In place of the term A.^ ^. - cpl(JJ^^J' it is convenient to substitute A. - cpiij, into 
eq.(2.220). 

Thus, the calculation of a rotor with variable pitch is acconplished by 
means of formiilas derived in this Subsection and differing only by the fact that 
ai and bi are not equal to cpi^j^^ and Ti(i(.) ^"d are found from eq.(2.22l) after 
determining cp^, . and cpl(^^ • 

The angles 9i(ic) and 9i(^n should be equal to the angles between the plane 

of the blade tips and the plane relative to which the rotor pitch is constant, 
i.e., the plane of the equivalent rotor. Consequently, 



<?•(«)= -\. 



= a, 






(2.222) 



Thus, without introducing the concept of an equivalent rotor we obtained /125 

131 



eqs. (2.220) and (2.221), after actually relating the quantities pertaining to 
the plane of the blade tips vjith their corresponding quantities of an equivalent 
rotor. 

The formulas derived in this Subsection yield the conditions of equivalence 
of a rotor vdth and -without flapping hinges: Rotors are "equivalent" if their 
angles of attack differ by a quantity equal to ai, and a rotor without flapping 
hinges has a coning angle and conponents of cyclic pitch change determinable by 
eq.(2.220) or eq.(2.22l). 

Here, it is assumed that the flapping hinges are located on the axis of 
rotation of the rotor or close to it and we can disregard the effect of second 
and higher harmonics of flapping on the aerodynamic characteristics of the rotor. 

The geometric meaning of the conditions of rotor equivalence is that, ipon 
satisfying these conditions, the position of the blades of both rotors relative 
to the velocity vector of the oncoming flow and their setting angles at all azi- 
muths are identical. It is obvious that, in this case, the thrust moment of the 
blade relative to the axis of rotation of a rotor without flapping hinges is 
equal to zero. 

If, for a helicopter with a rotor without flapping hinges, the cyclic varia- 
tion of rotor pitch for balancing the longitudinal and transverse moments is 
such that eqs. (2. 220) are not satisfied, then the aerodynamic characteristics 
of the rotor differ from those of a rotor with flapping hinges. For exanple, 
by creating a transverse moment by a lateral shift of the center of gravity of 
the helicopter toward the side of the advancing blade (ij/ = 90°), we can reduce 
the angles of attack of the blade sections at azimuth \|f = 270° and thus eliminate 
flow separations for a rotor without flapping hinges. 

25. General Expressions, for Determining the Components 
of Blade Pitch Change cpo> 9i > ^^'^ ^1 

In Subsections 23 and 24, we presented a method of calculating the aero- 
dynamic characteristics of a rotor with a blade pitch cyclically varying in the 
first harmonic 

?=9o — ?iCostJ- — tpiSin<|.. (2.223) 

Let us now derive fonnulas for determining the cciiiponents of blade pitch 
change cpo> 9x > ^^^ cp^ . 

The blade pitch established by the control units of the helicopter - control 
of the overall rotor pitch and inclination of the automatic pitch control mecha- 
nism - is represented in the form 

e = 9n-&iSirnJ>-62COS>p. (2.224) 

We assume that the design and working principle of the automatic pitch 
control are known to the reader [see, for example (Ref.l2)]. 

132 



In addition, the blade pitch of helicqaters is iisually changed during blade 
flapping, which is achieved by a special arrangement of blade tiorning levers and 
flapping hinges. The hubs of such rotors are called hubs with "flapping com:- 
pensator". Let us examine several schemes of hubs with a flapping conpensator: 
Cardanic and non-Cardanic hubs differing in control of blade rotation about the 
axial hinge, and also hubs with an offset and with a turned flapping hinge 
(Fig. 2.47). 



a.r 




Co ar 



Pm^P 




/126 



a.r 




Fig. 2. 47 Schematic Sketches of Rotor Hubs. 

a - Cardanic; b - Nonr- Cardanic ; c - With offset 

hinge; d - With turned hinge; v.h = Vertical 

hinge; a.r = Axis of rotation; 

h.h = Horizontal hinge. 

In the first scheme (a), the blade turning lever does not participate /127 
in moving the blade relative to the drag hinge but participates in others. In 
the thii^i scheme (c), the flapping hinge is located such that, in horizontal 
flying regimes, the blade axis is practicaUy perpendicular to the axis of the 
flapping hinge and goes through the middle between its bearings. 

In these schemes, the interdependence of setting angle and flapping angle 
of the blade is acconplished by displacement of the ball bearing of the blade 
lever A from the axis of the flapping hinge (n / O). In the fourth scheme of 
the hub (d), the interdependence of pitch and flapping angle is achieved by 
rotation of the axis of the flapping hinge. 

In all schemes, the blade is shown in a position inclined about the drag 
hinge through an angle Cq = lav • The flapping angle of the blade p is in a 
plane perpendiciilar to the plane of rotation and goes through the axis of the 
blade. Since the angle Cq is small, the angle of turn about the flapping hinge 
in the first three schemes can be considered equal to the flapping angle of the 



133 



blade: P^.h = 



P 



cos Co 



- p and, in the fourth scheme, as equal to 



K..-- 



cos (c, — Co) 



(2.225) 



We derive the forrmulas in the follovdng sequence: First we determine the 
mode of blade pitch change if the axial hinge had seized and the ball bearing 

of the blade-turning lever A was discon- 
nected from the rod of the automatic pitch 
control mechanism. This change in blade 
pitch, taking place without turning of the 
blade in the axial hinge, is called "kine- 
matic change of pitch" . We will denote it 

W /^ i n • 

We then determined the amount by which 
the blade is tiorned in the axial hinge, 
owing to the fact that the point A is con- 
nected by a rod with the automatic pitch 
control and cannot be displaced in flapping. 
This change of pitch is designated by /kpa.h . 
The overall change of pitch Acpo is equal 




to the sum of /kpi^jn and Zkipa 



Acp^ = Acpki„ + At? 



a.h 



(2.226) 



Fig. 2. 48 



Kinematic Change of 
Blade Pitch. 



The kinematic change of blade pitch in 
flapping is due to the blade axis being per- 
pendicular to the axis of the flapping hinge. 
Its derivation is clear from Fig. 2. 48. 
Point B, referring to the leading edge of 
the blade, during flapping of the blade is displaced relative to the plane of 
rotation by a greater amount than point B' referring to the trailing edge. Conr- 
sequently, the blade changes its angle relative to the plane of rotation: 



A?kin=-T- 



As 
b 



_ _i_sincolhA- — 



PtonCQ. 



(2.227) 



The derivation of /kpkin is illustrated also by the drawing shewn in 
Fig. 2.49. 

It is obvious that, when the flapping hinge rotates together with the blade 
during flapping of the blade relative to the drag hinge (hubs of Sikorsky heli- 
copters. Fig. 2. 50), there is no kinematic change of pitch. Such a change is 
virtually absent in the scheme shown in Fig. 2. 47c (hubs of Mil' helicopters) 
since, in horizontal flying regimes c 



hut 



A9|^.^ ^^Ptan(Co- 



Co; 



-O-O- 



(2.228) 



In the scheme of a hub with a tijirned flapping hinge (scheme d in Fig. 2. 47; 

such a scheme for the drag hinge is sometimes used for tail rotors of single- /128 
rotor helicopters), 

134 



A?ui,s=~PM<'i-Co). 



(2.229) 



Wow let us find Acpa.h • ^ point A in Fig. 2. 47 were not connected by a rod 
with the automatic pitch control mechanism, then during flapping of the blade 
it would be displaced relative to the axis of rotation by an amount As = nP for 
a Cardanic (universal) hub and As = [n +(lv.h - n)c| + I^CqIP » [n + t^,co]3 for 




Fig. 2-49 Kinematic Change of 
Blade Pitch. 



cw 




Fig. 2. 50 Schematic Sketch of Hub of 
Sikorsky Helicopter. 



a non-Cardanic (nonuniversal) hub. Since point A cannot have such a displace- 
ment, the blade in flapping turns about the axial hinge by an amount of 



As n a X a 



(2.230) 



(2.231) 



(2.232) 



for the universal hub and 

A9^^ = -(-p + '^o) P= -(tanai + Co)P 
for the nonuniversal hub. 

For a hub with an offset flapping hinge, we have 

A?a.V, -= - [tan =>! + (Cq - C^J] P « —tan OiP. 

For a hub with a turned hinge, Apa.h = 0. 

Thus, the total change of blade pitch during flapping motion is equal to: 
for a universal hub 

Acp^ = Acp^.„ + Atp^.v, = ( — txn o, +t»n Cq) P«s — (w o, — Cq) p; ( 2 . 233 ) 

for a nonuniversal hub and a hub with offset hinge 

A<p,==-tanai.p; (2.234) 

for a hub with turned hinge'"" 



-" For sinpllcity in Subsection 16, we take Cq = 0, Acpg = -P tan g^ 



135 



A9c=-tan(oi-Co)p«-(ta„Oi-Co)p. (2.235) 

The dependence of pitch on the angle P, in the general form, is esqjressed by 

A'ft=-*?. (2.236) 

where k is a coefficient of the flapping conpensator. 

The value of k is determined from eqs.(2.233) to (2.235). 

Oiir derived expressions for k do not take into account additional changes 
in the setting angle, such as those caused by inclination of the rod of the 
automatic pitch control mechanism, etc. Therefore, the quantity k should be 
corrected by measurements on a manufactiired hub or its model. This is deter- /129 

mined as the partial derivative k = _2x_ at an average blade pitch and blade 

angle of deflection relative to the drag hinge equal to Cq. 

For further confutations, the quantity k is conveniently represented as the 

tangent of some angle 6 : 



^(p(With consideration 



of elast ic twi st) 




Average elastic twist 
Vq per revolution 

Deformed 
blade 



Unde formed 
bl ade 



Ai=tftn8. 



(2.237) 



/a r' 



Fig. 2. 51 Rotor Pitch and Blade 
Twist, vdth Consideration of 
Average Elastic Twist per 
Revolution. 



The blade pitch is equal to the sxmi 
of the angle 9 established by the conr- 
trols and the angle bss} c ' 

9=00 — OiSiniJ. — ejcostp— ;fep. (2.238) 

Substituting into eq.( 2.238) the 
expression for P with an accuracy to the 
first harmonic, we obtain 



^ — (e„ — Aao) — (6 , — A6j) sin (jj — 
— (Oj — Afflj) cos ij). 



(2.239) 



Conparing this with eq.(2.223), we find 






(2.240) 



We recall that, in the calculations by the Glauert-Lock theory, the blade 
pitch is coiinted from the zero- lift angle of the profile: 



P^l.r.t- 



■«0 = 6n — ^^n — a„ 



(2.2400 



136 



For a rotor with a flapping conpensator, the equivalent rotor theoiy does 
not take into accoimt the change of blade pitch with respect to harmonics higher 
than the first, in view of the fact that this change is produced by higher har- 
monics of blade flapping. Higher harmonics of the change of pitch can be ac- 
counted for by specially derived forroulas. 

This also pertains to the average elastic twist over the blade with respect 
to higher harmonics and also to higher haraionics of the change of pitch due to 
elasticity of the automatic pitch control mechanism. The average elastic twist 
over the blade in the first harmonic Vi and Vn must be introduced into the 

* a V a V 

expressions for cpi and cpi . The average twist per revolution vq must be subdi- 
vided into average twist over the blade Vq and variable twist over the blade 

radius Vq - Vq ; the first is introduced into the expression for cpo, and the 

second is added to the geometric twist of the blade Acp (Fig. 2. 51) • 



Averaging of the elastic twist of the blade (this can be determined in 
flight tests or by calciilation; an estimate of the magnitude of twist can be 
made from the magnitude of the hinge moment of the blade) is carried out by 
means of the formulas 



/r^ 



j vor2dr 
_o 



1 





1 
K =3i•V^^^^«^>;=o.r• 



(2.2A.l) 



form 



With consideration of the elastic twist of the blade, eqs.(2-240) take the 






(2.242) 
(2.243) 
(2.24^) 



With the use of eqs.( 2.161) and (2.162), we find 



&v 



Solving this system relative to cpx and cpi we obtain 



?r 



1+A;2 



l+ft2 



1 +*2 



(2.245) 

137 



1+^2 1+^2 1 + ^2 • (2^246) 

26. Determination of Flapping _Coef ficients of 
Rotor with Flapping; Compensator 

After substituting eqs.(2.245) and (2.246) into eqs.(2.l6l) and (2.I62), we 
obtain the following e^qjressions for the flapping coefficients ax and bi : 

^ _ ai, + ^»i -61+^6; , -^'av+^ "^'av . (2.247) 

' 1 + A2 "T" 1+^2 '^ 1 + *2 

^ »,,-fe^, e, + A6i , %v+^^iav (2.248) 

1+^2 ' 1 + /fe2 ' l+k2 

The first addends on the right-hand side of eqs.(2.247) and (2.248) deter- 
mine the flapping motions of a rotor having a flapping confjensator td-th an unde- 
flected automatic pitch control mechanism G2 =61 =0 and without consideration 
of elastic twist. For sinplification of the formulas we write them in the form 

- _ aie + fe^ie . (2.249) 

U-i — ■ , 

' 1 + ft2 ' 

*'=TT^- (2.250) 

The presence in the kinematic scheme of the hub of a flapping conpensator /I3I 
greatly affects the flapping coefficients of the rotor. Upon an increase of k 
the inclination of the axis of the rotor cone to the side of the advancing blade 

_ bi - 

(i|f = 90°) bi decreases. When k = ° we have b^ = 0, and with a fiorther in- 



1. 



crease of k, bi becomes negative, i.e.- the axis of the cone is incHned to the 
side of the retreating blade (^ = 270°). 

The backward inclination of the axis of the rotor cone aii ipon an increase 
in k varies differently. At a small value of the coefficient of the flapping 

conpensator (k < -r-^ — ), the axis of the rotor cone is still inclined toward the 

advancing blade and the setting angle of the blade increases at azimuth i|f = 90° 
(A9k = -kp) and decreases at azimuth \|f = 270°. Therefore, the coefficient ai 

increases. At k = —-ia— , bi = and the setting angle at azimuth t = 90° and 

t = 270° will not change. Thus, ai = ai . On further increase in k, the angle 

b^ becomes negative and the setting angle at azimuth i|f = 90° decreases, whereas 
at azimuth i|f = 270° it increases. Therefore, the coefficient aii decreases. 

138 



For a single-rotor helicopter, the tail rotor has no automatic pitch con- 
trol so that the blade flapping is determined by the quantities ag, a^, hi (the 
elastic twist of the tail rotors being small) . 

The maxom-um flapping angle of the blade in this case is equal to 

P.na. = «o+AP=«o + l^afT^5 = ao+^'^!^^ (2-251) 

Equations (2.249) - (2.251) show that the flapping conpensator decreases 
the magnitude of the variable portion of the flapping motiori^^ Ap and changes the 

azimuth at which the flapping angle has an extreme value tan \|rR = _-^ . Both 

m 1 n ai 

the first and second factor may be of significance for a single-rotor heUcqpter: 
the first decreases the variable loads on the blade of the tail rotor and the 
second changes the gap between the blades of the tail rotor and the tail boom. 

For helicopters of side-by-side configuration, the gap between the blades 
and the fuselage is deteraiined mainly by the quantity bi . It is expedient to 
select a kinematic scheme of the automatic pitch control mechanism such that a 
deflection of this mechanism will not influence the quantity bi (usually only 
course control of a helicopter is acconplished by the automatic pitch control, 
i.e., change in ai). Consequently, for helicopters of side-by-side configura- 
tion we must take into account, when selecting the magnitude of the flapping 
conpensator and disregarding the elasticity of the automatic pitch control and 
of the blades bi = bi, that the quantity b^ should be higher or lower depending 
on the direction of rotation of the rotors. 

For coaxial helicopters the 'gap between the blades of the ipper and lower 
rotors depends on the quantity b^ (because of the mutual interference of the 
rotors, the operating conditions of the ijpper and lower rotors are not the same 
so that also the difference of the coefficients ai of the rotors has an influ- 
ence on the gap). As we see from Fig. 2. 52 this gap is 

A = // + /?P.f. -/?Pu-.=^-^(^.y. +*>,...>• 

It is obvious that to increase the gap we must select a magnitude of the /132 
coefficient of the flapping conpensator such that b^^ ~ ^ii =0. 

The value of b^ of single-rotor and fore-and-aft helicopters and the value 
of a^ of helicopters of any configuration are deteraiined by balancing the heli- 
copter. The pilot, deflecting the control stick and acting on ai and bi , 



^;- R.A.Mikheyev determined that Ag decreases somewhat less than Vl + k^-fold, 
since tail, rotors with k = and k ^ should be considered at an identical 
angle of attack and since of, of a rotor with k ^ is less than a^ of a rotor 
with k = 0; due to this fact, the quantity ai^ of the former, at equal thrust, 
is greater than that of the latter. 



139 



'pp er rotor 




Og-bt 



Fig. 2. 52 For Determining 

the Gap between Rotors of 

Coaxial Helicopters. 



establishes ai, bi, h, and s in such a manner 
that the helicopter is in balance. However, at 
the required values of ai or bi the quantities 
ai and bi depend on aii and bi ; consequent- 

oon oon ^ 

ly, the flapping conpensator influences the de- 
flection of the helicopter controls in flight, 
i.e., its "balancing curves". 



27* Determination of the Components of Blade 
Pitch Change cpi and ^i after Deflection of 



the Automatic Pitch Control 



The second addends on the right-hand side 
of eqs.( 2.245) and (2.246) determine the incre- 
ment of the conponents of cyclic pitch change and the flapping coefficients of 
a rotor with a flapping conpensator, after manipulation of the automatic pitch 
control. They represent the change in position of the blade cone and the direc- 
tions of forces and moments relative to the rotor shaft when the helicopter 



controls are manipulated. 



¥e denote these by cpi and cpi 

" "'• oon "■'•CO n 



or a^ 



and bi 



^' con ' con 



— Bi + yfeBa 
1 +ft2 



= *, = 



6.; + feO] 



"" 1+ ft2 • 



(2.252) 
(2.253) 



Let us now establish the relation between the angles of inclination of the 
automatic pitch control and the magnitude of the angles 9i and Gs . 

Figure 2.53 shows a diagram of the hub and automatic pitch control in top 
view along the rotor shaft. The arrangement of the flapping hinges is not shown, 
since this has no influence on our derivations (only the expression for k de- 
pends on it) . The segments AA.' are projections of the inclined rods of the 
automatic- pitch controls; point A is the coipling of the rod with the blade 
turning lever, while point A' is the coi:pling of the rod with the automatic 
pitch control itself. 

The rotor blades are shown in positions at which Acpeon = ^s (if = 180°) and 
Acfbon = 01 ('I' = 270°); here the hub and automatic pitch control are turned, rela- 
tive to the longitudinal axis of the helicopter, through the angles ^^^^ = 
= 180° + Co and %^^y, = 270° + Co- 

The angles of turn of the automatic pitch control will de denoted by the 
letters k and Tj, where k is the tiirn mainly causing deflection of the blade cone 
in the longitudinal plane of the helicopter and Tj in the transverse plane. 

Let the automatic pitch control be deflected through an angle k relative to 
the axis OO' located at an angle Ailfoon "to the transverse plane of the helicopter. 



140 



ili=t80 
AfBi 



if)=270'> 




\ Longitudinal axis 
\0 t of helicopter 



Fig. 2. 53 Kinematic Diagram of Hub and Automatic 
Pitch Control. 



7133 



Considering that the vertical displacements of the points A and A' are 
equal to (S,^ = Sj^), we find 



6^^> = ^=*^?5:f< 



02 = — =*-pfsin (a.i-Atcon -Co)=Z>2x; 



s!") 



.R 



= -x--^COs(a„-A<l'„n -Co) = -£'i'<- 



(2.254) 



Similar expressions are obtained on deflection of the automatic pitch con- 
trol relative to the axis 00" through an angle T] 






(2.255) 



Let us determine the increments of the flapping coefficients of the rotor 
when the automatic pitch control is deflected through the angles k and T), sub- 
stituting the obtained relations into eqs.(2.252) and (2.253; 



,,, ^ /?a.p cos(oii — A^/^^^ — co) + fesin(aii-Atc,n — gp) _ 



l+*2 



«a.p 



a\ 



(1) 



=x ^ cos 8 cos (0,1 - A-P^n - Co - 8); 
»b 

- Ti ^ cos 8 sin (0,1 - .^ - Co - 8); 



* '?.n= * y* ^°^ ^ ^^" ("" -' 1*000 - '^o - 8); 



(2.256) 

(2.257) 
(2.258) 



141 



•c»n ' l^ ^ " -l-con (2.259) 

Usually, the angles are so selected that cos (an - Ailfcon - co - S) > 
> sin (ctxi - Moon ~ Co - S)» consequently, the deflection of the automatic 
pitch control through an angle h mainly causes a change in the coefficient a^, /134 
whereas a deflection through an angle T] will change the coefficient hi . The 

product I ^^ cos 6 cos (on - Moon - Co - S)» which is dependent on the kine- 
matic scheme of the hub, constitutes a relation between the longitudinal in- 
clination of the axis of the rotor cone and the angle h . This is denoted by D^ : 

D^=^coshcos{o^^~^^^,„-Co-^). (2.260) 

The product ^ ^^ ' cos 6 sin (cth - Ai|/con - ^o ~ ^) characterizes the in- 
clination of the axis of the rotor cone in a lateral direction. TMs is denoted 
by Dg: 

Dj^^:Pros«sin(a„-Atc,„-^o-8)- (2.261) 

'b 

The value of the coefficients D^ and D^ can be refined by testing the full- 
scale hub or its model. For this, the blades are set in an azimuthal position 
(shown in Fig. 2. 53) and relative to the flapping hinge at an angle P = Pav = ^-o. 
After deflecting the automatic pitch control through an angle k, the increments 
of the setting angles, i.e., the angles Bg and 9i , are measured. 

The values of D^ , Dg, D^ , and Dg are found from the e:!^ressions 

X 

' 1 + A2 X 1 + ft2 

^ ^ Dj — kPi ^ 1 62 + Ml 
^ l+k2 X l+k2 ' 

The quantity k is also found from tests (see Subsect.25). 

Thus, 

?icoa=-^'co„ = -'^i''+^2Ti; (2.262) 

^'con = \.n='^l^ + ^2-''- (2.263) 

142 



Helicopter designers often acconplish the kinematics of the automatic pitch 
control in which D^ = 0. This is done so that, with a longitudinal deflection 
of the control stick causing inclination of the automatic pitch control by an 
angle k, only the coefficient ai is changed, i.e., so that the resultant force 
is deflected strictly in the longitudinal plane of the helicopter. This creates 
a moment relative to the center of gravity of the helicopter, also acting in the 
longitudinal plane of the craft. However, the motion of helicopters is so in- 
terconnected in all directions that there is no sense to rigorously insist on 
the condition of coincidence of the directions of action of the moment and de- 
flection of the control stick. 



For helicopters of side-lDy-side and coaxial configurations, for which 
special demands are made on the quantity bi, the coefficient Dg should he equal 
to zero so that hi does not change when the automatic pitch control is deflected 
longitudinally foCTsrard or backward. 



We see from eq. (2-261) that Dg = 0, when 



(2.264) 



If, in the kinematic scheme of the hub and automatic pitch control, a value 

of axi - Co - S 7^ is obtained, then the 
plane of inclination of the cone axis will 
not coincide with the plane of inclination 
of the automatic pitch control but will lead 
the plane of inclination of the automatic 
pitch control by an angle of 



Longitudinal axi 




!:=c 



-c„ — 8. 



(2.265) 



Fig. 2. 54 Position of Blade at 
Instant of Maximimi Pitch Change 
(Hub without Flapping Gompenr- 
sator) . 



Let us explain the derivation of the 
lead angle, assimiing - for sinplicity - that 
the coefficient of the flapping conpensator 
is equal to zero (6 = O). In this case, the 
cyclic change of the setting angle is pro- 
duced exclusively by inclination of the auto- 
matic pitch control. 



Figure 2.54 shows a blade in a position 
at which its setting angle has a maximum 
value since the point A' of the rod, connect- 
ing the blade turning lever with the auto- 
matic pitch control, lies in the plane of 
inclination of the automatic pitch control. The plane of inclination of the 
cone axis is perpendicular to the blade position. We see from Fig. 2. 54 that 
C, = a-ii - Co . The angle C, is nonzero since the mechanism changing the pitch is 
designed such that, for a maximum change of blade pitch at some azimuth t> the 
automatic pitch control will be deflected at an azimuth differing by an angle 
90° - an + Co . If CTii 7^ Co, then the plane of inclination of the automatic 
pitch control will not coincide with that of the cone axis . 

At 6 ^ 0, the cyclic change of the setting angle is not only directly due 

143 



to deflection of the automatic pitch control but also to the fact that the 
change of flapping, caused by deflection of the mechanism, in turn changes the 
angle of blade setting. Here, the azimuth of the maximum total change of setting 
angle lags by angle 6 = tan~''"k behind the azimuth of the maximum change of set- 
ting angle due directly to inclination of the automatic pitch control. There- 
fore, at 6 f^ the lead angle is determined by eq.(2.265;. 

For the plane of inclination of the cone axis to coincide with the longi- 
tudinal or with the transverse plane of the helicopter (so-called "independent" 
control), the axes of inclination of the automatic pitch control 00' and 00" 
should be turned to the longitudinal and transverse planes of the helicopter 
through an angle M/aon = £♦ 

It follows from eq.(2.26o) that, with "independent" control, the coeffi- 
cient will be 

^'=lf''''- (2.266) 

b 

If for a hub cq = the rod of the automatic pitch control is vertical, then 

Z126 

^=—1- (2.267) 

'b cos an 

and 

n cos 8 / . N 

^'=7^- (2.268) 

28. Sequence of Aerodynamic Calculation of .a Rotor 
with Variable Pitch 

Thus, the eDspressions for the conponents of cyclic pitch change are written 
in the form 

— — — - _ v"i +ftv, (2.269) 



Vi^-DiX-j-D^n-kbi +v, =-DiX+D2Ti-Afti 






l+*2 (2.270) 



For brevity, the last addends of eqs.(2.269) and (2.270) are omitted in 
what follows. 

Let us now. derive the expressions for determining the coefficients h and s: 

= h, + tkb,-^ta, . ^^'^^^^ 

* ' 'con ' 

s=5,+%=s.-/Aa, + ^(DiTi+D2x)= (2.272) 

=s^ — ikai-]-ib . 

con 
IW- 



The flapping coefficients of the rotor are 






(2.273) 
(2.274) 



For helicopter flight with sideslip, eqs.(2.269) - (2.274) should be cor- 
rected. For exanple (Fig. 2. 55) » 



fe, =6, + (DiTi + D^x) cos p, j+ (Dix - DjTi) sin p^ ,, 



(2.275) 
(2.276) 



since the inclination of the cone axis and of the aerodynamic force produced ■by- 
deflection of the automatic pitch control is determined along axes related with 

the helicopter regardless of the direction of the 
velocity -vector. The angles a^, a^, bi, bx, just 
as the forces H and S, are the angles and forces 
along axes fixed with respect to the direction of 
the velocity vector. 

If we represent h, and s^ in terms of conpo- 
nents lying in the plane of the blade tips 
[eqs.(2.207) and (2-208)], we obtain 

A = A(K) + ^ai = A(K) + /ai + /(DiX-D2Ti); (2.277) 
s=S(K) + /&,=S(K)+<&, + /(£>,Tl+^2'')- (2.278) 

Equations (2.269) and (2.270) show that_, for /137 
a rotor with a flapping conpensator, cpx ^-^-d cpi de- 
pend not only on the angles of deflection of the 
automatic pitch control but also on the flapping 
coefficients of the equivalent rotor. This sub- 
stantially conplicates the calculation of a rotor 
with a flapping conpensator since, in determining 
the initial data for calculating the equivalent 
rotor \g, 9, |j., it is necessary to know the coeffi- 




Fig.2.55 Deconposition 
of ai and b, into 

•^con con 

Velocity Axes during 
Helicopter FUght with 
Sideslip . 



cients a^ , b^ 



However, when any five quanti- 



ties characterizing the operation of a rotor with variable pitch are prescribed, 
it is always possible to select a calculation sequence (sometimes pre-assigning 
several values of X^ oi* 9 and constructing auxiliary graphs) which will contain 
all coefficients of forces, moments, and flapping of the rotor. 

Let us give a typical exanple. At given t (rotor thrust approximately 
equal to helicopter weight), iJ,, %, k, and T], the aerodynamic calculation se- 
quence for the rotor can be as follows: Assigning various values of \^ , the 
expression obtained from eqs.(2.157)"" and (2.242) 

* We recall that cpo entered the formulas in Subsections 3-24. For sinplicity, 
the subscript "0" of cp was omitted. 

145 



<Pe=^0 = - ^ , 

1 +-^(1+1^2) 
4 

will yield cpoj after which eq.(2.50) will give t^ = t. 

After deterndning, either by trial and error or graphically, the values of 
Xe and cpo at which t is equal to the prescribed value, eq.(2.40) will furnish 
ao, ai^ and bi^ ._ Then, ^qs.(2.^9), (2.250), (2.269), and (2.270) will be used 
for determining ai, bi, cpi, and cpi . We now have all data necessary for calcu- 
lating the characteristics of the equivalent rotor and their conversion in the 
axis of a rotor with variable pitch. To detemdne h, s, and m^ we can also use 
^formulas derived in Subsection 24, obtaining - in the above-described sequence - 
q'i(k) > 9i(k) > ^^d M)!.) ^^°^ eqs.(2.222). 

The aerodynamic calculation and the calculation of helicopter balancing are 
performed in the same sequence. 

As shown in Subsection 23, the aerodynamic characteristics of a rotor in a 
velocity coordinate system - t^ = f(ty, m^) at constant values of |j, and Mq - do 
not depend on cpi and ^i ; therefore, the conputation can be performed from the 
characteristics of a rotor with constant pitch: t^ = f(ty^ , m^ ) for the same 
values of |j, (p. = V) and Mq . From the aerodynamic calculation, we obtain the co- 
efficients tjj and ty ; at any value of cpi and cpi in a given flying regime, the 

characteristics of the equivalent rotor" will not change and will correspond to 
the found values of |j,, tx , and ty . Thus, as a result of the aerodynamic 
calculation, we will obtain all characteristics of the equivalent rotor. /138 
After this, we calculate 9o = cpo + kao, % , bi and, from eqs.(2.27l), (2.272) 
or (2.277), (2.278), the components h and s which do not depend on a.^aon ^^^ 
bx . From the condition of helicopter balancing, i.e., from the condition of 
equating to zero the longitudinal and transverse moments, we find ax and bx 
and follow this by calculating k , J\, ax , bx , h, and s . 

Section 3. Momentum Theory of Rotor 

In the momentum theory of a rotor, the aerodynamic forces and the power re- 
quired by the rotor are found by applying general theorems of mechanics to the 
flow around the rotor. 

This theory is used in approximate calculations in which both the induced 



"- The characteristics of the equivalent rotor for ar^r cpx are the same when as- 
suming that the parasite drag of the fuselage (and also the angle of attack of 
the wing of a winged helicopter) does not depend on cpx and is determined as some 
average value (9x)av[Q' = c^s - (9iav)]« Usually, this assunption is valid. If 
it is not, it will be necessary to perform second-approximation calculations 
from the value of ^x obtained from the balancing calculation. 

146 



and profile power of the rotor are determined on the basis of sinplifying as- 
sunptions or from precalculated graphs.. In this case, there is no need to de- 
termine the angles of attack and elementary aerodynamic forces in each blade 
section, a fact responsible for the sinplicity of the formulas. 

In the momentum theory, the con^jonents of the aerodynamic forces of the 

rotor along the velocity of flight (drag) 
and normal to it (lift) are determined, 
which makes this theory convenient for 
use in helicopter calculations. 



1. Theory of an Ideal Helicopter Rotor 




When creating lift and drag (or 
propulsive force), the rotor thrusts an 
air mass downward and forward (or back- 
ward) . 



Glauert postulated that the rotor 
acts on an air mass passing through the 
area of a circle placed normal to the 
flow incident on the rotor. The diameter 
of the circle was to be equal to the 
diameter of the rotor (Fig. 2. 56). This 
postulate is based on the fact that the 
same flow boundaries are selected both 
for the propeller and for the wing, with 
uniform induced velocity distribution. 
For the propeller, this is entirely ob- 
vious since the flow boundary is deter- 
mined by the area swept by the blades; 
for the wing, the possibility of select- 
ing such a flow boundary is given by the 
vortex theory. Recently developed vortex 
theories of a rotor rather- accurately /I39 
confirm the correctness of Glauert 's 
hypothesis concerning the air mass participating in the generation of the aero- 
dynamic forces of a rotor. 

In the ideal rotor theoiy (Ref .21), it is postulated that the air flows at 
the same velocity over the entire area of the circle: The air stream does not 
mix with the surrounding air, so that it is proposed that the air is an inviscid 
fluid. Fiorthermore, it is assumed that profile losses of power and vorticity of 
the stream are absent. 



V^-V-Vx^ 



Pig. 2. 56 



Model of Airflow around 
Rotor . 



A model of the airflow and its velocity coirponents in three sections - far 
ipstream of the rotor (section O-O), along the rotor axis (section 1-1), and far 
downstream of the rotor (section 2-2) - are shown in Fig. 2. 56. The induced ve- 
locity corresponding to the rotor lift Y is denoted by the vector Vy, while the 

induced velocity corresponding to the rotor drag X, is represented by the 
vector Vx ( Vy W^f, v, \\^) and the velocity of the undisturbed flow, by V. All 

147 



vectors are shown for positive direction, with the subscripts corresponding to 
the section number. 

From the theorem of moment of moment\mi follow the relations 



where m is the air mass flowing per second through the section 0-0, 1-1, or 2-2. 
The variation in kinetic energy of the per-second air mass is 

E=E,-Eo = -^m{Vl+vl~V^). (3.2) 

Equating eq.(3.2) to the expression of energy inparted to the air by the 
rotor in unit time 

(3.3) 
we obtain 

After transfonnation, this es^jression is reduced to the form 

which indicates that, in the examined flow, the following conditions are satis- 
fied: 



^i/. =-77 '"y 



(3.4) 



2 
and 

T'^^'- (3.5) 



■^•^. =^r '"•'.• 



These relations show that the induced velocities in the rotor plane are 
one half those far downstream of the rotor. 

The power supplied to an ideal rotor is e^qjended only for creating kinetic 
energy of the flow and thus is equal to it. Making use of eq.(3.3), we have 

75N=V-Vy,-XiV~VjcJ. (3.6) 

The weight rate of flow of air m per second is equal to the product of the 
mass density and the volume rate of flow of air per second: 

m=QFV', (3.7) 

where V' is the resultant of the velocities V, Vy and v^ : 

148 



V' = YV]+ vl=y(V-v,,f-\-vl. 



(3.8) 



To account for the so-called tip losses of the rotor, the following /1Z^0 
method is used: In calculating the forces and induced velocities, the air mass 
flowing through an effective area F^tt = B^F smaller than the area swept by the 
blades is introduced. Usually, for forward flight regimes we take B^ = 
= 0.94 - 0.96 [the effective radius of the rotor R^tt = BR = (0.97 - 0.98)R]. 
The power is calculated on the basis of the mass of air flowing through the 
actual area. Therefore, by means of eqs.(3.l) and (3.4) to (3.8), with consid- 
eration of tip losses, we obtain 



75N=:^iyv,-X(V-v„)]. 



(3.9) 
(3.10) 
(3.11) 



Let us now change over to dimensionless quantities - coefficients of forces, 
moments (power), and velocity: 






m 



\—^[(^y'^y-Cx{V—vX 



(3.12) 
(3.13) 
(3.1!j-) 



In these expressions we omitted the subscripts since, from now on, we will 

be concerned only with the velocities in 
section 1-1 and, for sinplicity, will not 
give them a subscript. 

Equations (3.12) to (3.14) describe the 
general case where any aerodynamic system 
creates lift and drag (or propulsion) and 
consumes or yields power. Therefore, these 
expressions also are valid for a propeller 
and for a wing. For a propeller, Cy = Vy = 
= must be substituted into eqs.(3.l2) to 
(3.14) and for a wing which does not inject 
engine power into the flow, m^ = 0. (For a 
wing, one usually takes "v, < V, Vy < V. ) 




Fig. 2. 57 Deconposition of the 
Resultant Aerodynamic Force of 
a Rotor into Velocity and Body- 
Fixed Axes. 



In the rotor theory, it is conventional 
to use quantities in rotor-fixed coordinates 
Cj, C^ , [J*, A., etc. 



Introduction of the angle of attack of the rotor yields the following rela- 
tions between the coefficients of forces and velocities, in different coordinate 
systems (Figs. 2. 57 and 2.58): 

li..9 



Cj-^Cy COS a-\-Cx sin a; 

V( = "Wy cos a -f "Ux sin a; 
Cu=Cx cos a — Cy sin a; 



■Wft = 1)^ cos a — -Wj, sin a ; 
F'' = (Fcosa-^J + (F sina -^^j'^^a+X^; 
\i-^V zosa~v^\ 
X=: Vsina — z;^. 



(3.15) 
(3.16) 
(3.17) 

(3.18) 
(3.19) 
(3.20) 
(3.21) 



Substituting eq.(3.l2) and (3.13) into eqs.(3.15) and (3.17) we obtain 
formulas for the coefficients Cj and C^ : 









m 



t=— -^(^7->- + C//P')- 



iS2 



(3.22) 
(3.23) 
{3.2h) 

(3.25) 



Let us now study the velocity polygon of a rotor and derive a number of 

additional relations facilitating 
the calculation of rotor charac- 
teristics. 

The velocity and force poly- 
gons are shown in Figs. 2. 58 and 
2.59. The latter diagram, as a 
supplement to Fig. 2. 58, shows the 
vector of the restiltant induced 
velocity 




--'Vy-\r'"x='"f\- '"h 



Fig. 2. 58 Velocity Polygon of Rotor. 



as well as the angles C a-nd 5, as 
follows : 

C, - angle between the ve- 
locity of the undis- 
turbed stream (flying 
speed) and resioltant 
velocity in the rotor region, £ > at Cy > 0; 
6 - angle between the normal to the velocity of the undisturbed stream 
and resultant aerodynamic force of rotor. 



150 



Since Cr || u, the angle between the vectors Vy and u vdll also be equal to 



6. 



The angles C and 5 and the flying speed V conpletely determine the velocity 
polygon. To determine the velocity polygon in terms of vectors in a fixed co- 
ordinate system, one more quantity 
must be known, such as - for ex- /1^2 
anple - the angle of attack of the 
rotor a . 

Let us write out the main re- 
lations between forces and veloci- 
ties in the velocity polygon: 




£iL=^=Un8; 



Cy 






Fig. 2. 59 Velocity Polygon of Rotor. 



sin C cos S cos (C — -8) 

Cy = C/j COS 8, 



(3.26) 

(3.27) 
(3.28) 
(3.29) 
(3.30) 



'Uy = acos8, 1 
•u_j = asin8. J 

Using these relations, eqs.(3.l2) - (3.14) can be written in another form 



m 



Cy= AB^v.V ^4BW' ^'"^^"^'^ ; 
^ ^ cos2(C — 8) 

Cj^ = Cy 't«n8; 



x(^-g)='^^^-(^-^) = ^ ^^^-(^- 



8). 



(3.31) 
(3.32) 

(3.33) 



Equations (3.31) to (3.33) are of interest in that the two independent _ 
variables C and 6 correlate the coefficients Cy, Cx, and mt at ar^ value of V, 
which is a consequence of the similarity of the velocity polygons in regimes in 



C C 
which the ratios -=^ and -~- are equal. 
yS V^ 



Having assigned a series of values to the angles C, and 6, we can find 



/Mik 



the ratios 



m 



B^ 



2 * B^V^ * 7^ 



i- and construct a graph for their correlation. 



151 



U1 
JO 



# 



?J^##-^^^^^^^-^ 



^:'„«^L£ 




-0.10 —0.05 



0.03 „ , 
-O.Oit 

-'■''-00 

0.13 ";f 

0.1S 



Fig. 2. 60 Interrelation of Coefficients of lift. Propulsive Force, 

Cy 



and Torque of an Ideal Rotor 



SttS 



Wl' 



< 1.0 



E 



At Cy = (C = O), eqs.(3.3l) to (3.33) are not applicable; therefore, in 
constructing the graph, we used eqs.(3'13) and (3.14) transformed into 



Cx _ , 



i73 






B2V2 



(3.34) 

(3-35) 



Such graphs (Figs.2.60 and 2-6l) are convenient for solving problems of 
aerodynamic design. 



~= 0.5 m ^ :S - 4 -5 -6 J 




Fig. 2. 61 Interrelation of Coefficients of lift. Propulsive 

Cy 



Force, and Torque of an Ideal Rotor 



B^'^P 



= 1.0 



- 12.0). 



To determine the quantities entering the velocity polygon of a rotor, we 
must know - in addition to V - angles C and 6 . The angle £ is determined rather 
accurately by means of graphs (see Figs. 2. 60 and 2.6l), while the angle 6 is de- 

n 

termined from Cy and Cxt 6 = tan" " . 



153 



At small C and 6 (large V, small Cy ) eqs.(3.3l) to (3.33) are sinplified. /1^5 
Actually, we have 



Cy=452K=C; 



(3.36) 



Substituting the first two equalities into the third, we obtain 



m 



t— — (CyVt:. - CyVh)^^ ~ J- CvF. 

^82"-^ y ' 4B4K 52 '"^'^* 



(3.37) 



At small angles of attack 






^ C;,K. 



(3.38) 



The e^^ression for m^ is generally used in this form in aerodynamic calcu- 
lations of a helicopter in fljring regimes at V s 0.15. 

To calculate fljd-ng regimes with large £ (at small V), which are not 
covered by the graphs shown in Figs. 2.60 and 2.61, we must use eqs.(3.l2) to 
(3.14). Substituting into eq.(3.13) 



- - Cy 

^ Cx 



(3.39) 



and transforming, we obtain 



/Cx\2_2^ 



\4B2 , 



)=f [(^-^J+1/ (v^-^.r+ 



4B2 J 



(3.40) 



This expression permits constructing graphs for the aerodynamic character- 
istics of a rotor Cx = f(mt')> ^or any selected values of V and Cy: Assigning 
Vx, we find Cx, then Vy, and finally Ef, from eq.(3.14) (the sign of Cx coincides 
with the sign of v^ ) . 

To calculate the characteristics of an ideal rotor, we can also use the 
following expressions: 



C^ = 4B^uV' = 4B^u ]/(l/-usin8)^+(«cos8)^ = 



^AB'^u]/ 1/2 -a (21/ sin 8- u). 
Let us then transform eq.(3.4l) into 



(3.41) 



154 



1 = 



4B2 



Cr 
4B2 



■/■ 



Cr 
4B2 



2sin8 



/ 



Cr 
4B2 



(3.42) 



from where it is obvious that we can construct the graph of the dependence of 



u = 



V. 



on V 



i- 



and 6 . Such a graph, borrowed from another publi- 



4B^ ' 4B^ 

cation (Ref .2), is shown in Pig. 2. 62. In this diagram, the broken line is the 
approximate curve which can be used as basis for calculating the vortex-ring 
state at 6 = +60° and 6 = +90°-; for this curve, the ideal rotor theory does not 
hold. 

A determination of the induced velocity of the rotor is of independent 
interest . 

If the angles C a^nd 6 are known, the ratio of the induced velocity to /146 

the flying speed can be determined by expres- 
sions derived from eqs.(3-26) and (3.27): 
u 
2.0 

19 
1.8 

n 

1.6 
1.5 

t.'i 
I.J 
1.2 

1.1 

1.0 
0.9 
0.8 
0.7 
OS 
0.5 
0.1 
0.3 
0.2 
0.1 











/ 












1 


-_i 


J 

/ 
/y 

1/ 




\ 
1 


1 


- 


— 


















- 




















































y 


S-^QO" 
^60° 
^30° 
1^15° 
^^ 

^n<> 








^ 


1 


V 


\ 


-^ 










\ 




^ 








a 




-^ 










v3^ 










% 


\\Z^-60° 














^1— 


>}=-go- 














^ 


1^ 












— 


— 








^ 
















'Ni 


^ 


1^ 


_^ 





































V 


sin C cos B 1 

cos(C-B) 1 ' 
+ tan8 
tonC 


(3.43) 




i;^ = 'Uj,ta>i8. 


{3.kk) 



If the angle of attack of the rotor is 
known, the velocities Vt and Vjj can be ob- 
.tained from eqs.(3.l6) and (3.18). 

In the operating condition Ox = v^ = 0, 
the velocity Vy is determined from the ex- 
pression: 



"'-"/i/t+Ij^-T-t- (3- 



45) 



At small angles C a-nd 5, the relation 
"v-y. < "Sf ±s satisfied, so that 



Cy 



1.0 



2.0 3.0 1.0 5.0 V 



or 



Fig. 2. 62 Induced Velocity u 
as a Function of V and 5 . 



1),= 



Ct . 

4B2 ]/ 1/2^.^2 

Cx 

4B2 Yv^ +~vl 



(3.46) 

(3.460 
(3.47) 



155 



At V ^ 0.15 and at small angles of attack (more acciirately, if V ^ 2.0 and 
at small 8 where - in conformity -with Fig. 2. 62 - u « -s=-) the sinplified expres- 
sions widely used in the calculations will rather accurately yield the induced 
velocities 

- - Cy - Ct 

■u = z;„ = — s.- — 1), - - — 



4B2V 4B2(. (3^^g) 



(3.49) 

In these cases, the induced velocity is denoted by the letter v without 
subscript, for sinpHcity. 

2. Derivation gX t he Eajression for the Torcpae Coefficient 
of a Real Rotor 

Equations (3.1^) and (3.37) were derived above for determining the torque 
coefficient of an ideal rotor, when considering the rotor as an active disk in- 
fluencing its own circumflow. These e^qjressions are interesting in that m^ is 
represented as an explicit function of the coefficients_of Hft and propulsive 
force Cy and Cx . In the same foiro, the expression for m,. can be derived also /147 
for a real rotor. This derivation was originated by L.S.Vil'dgrube. The ob- 
tained equation is valid for nonuniform induced velocity distribution over the 
rotor disk and takes into account the forces of the profile drag of the blades. 

As is known from the classical theory (Sects. 2 and 4), the conponents of 
the dimensionless velocity of flow past the blade sections (at l-^^^ =0, 
cos p « 1) are equal to 

t7,=-^=r + 7cosasin<|-; (3.50) 

(7v = -^=Fsin a-ti — Fcosasln8cost>-r -^. (3.51) 

^ <»R ^ d^ 

The con53onents of the aerodynamic forces located in a plane perpendicular 
to the blade axis are expressed by the equations 

dt ={Cy cos ^ -\-c^pSinO)U''bdr', (3-52) 

rf9 = (c^pCOs«)-CySin«))t7^6afr. (3*53) 

Substituting, into eqs.(3.52) and (3.53), the e3?)ressions 

sin 0) = -^andcos© =-^^ , 

X.J- • U u 

we obtaxn 

dt=={r.p^^c^Jjy)Ubd'r; (3.54) 

156 



dq = {c,^U^-c^U,)Ubdr. (3.55) 

Solving eq.(3«54) relative to CyUbdr 

^ r;Tw " <" C:cfiyVldr 

and substituting this into eq.(3.55)» we obtain 

dq=c,i^,-\-^Ubdr-^Jt = c,^^%d7-dt -^. (3.56) 

The elementary torque of the rotor is 

dmt = dqr. (3.5?) 

From this, after substituting dq from eq.(3»56) and r from eq.(3.50) we obtain 

dm.^ = c^p^Fdr—dtUy — dqVcosas\n<\K (3.58) 

For further transformations of eq.(3»58) we use eq.(3.5l) and the e^qares- 
sions for the elementary longitudinal and propulsive forces: 

dh=:dqsin<if — dtsin'^cos'!f; (3-59) 

dt^=:dhcosa-\-dtsina. (3.6o) 



As a result we obtain /1Z(,8 

Li.VV Oil 

4Z ''P^ 



dmt = c^pU^bdr-\- dtv-\-dtr -^ — dtV s\na+dts\n pcos4>\/cosa- 



dqsXn'iiV cosa = c^pU^bdr-\-dtv-\-dtr-^— rf/Vsina— (3-61) 



d^ 
dl 
d^ 

We then integrate the elemental^ torques with respect to the rotor area 



~dhVcosa = c^/J^bdr-\-dtv-\-d{r-^-dt^V. 



2t 1 2x I 2it 1 2it 1 2ic I 

ntf. = frft r rfm^ = r f c^^'^dFd^ -{-U'Sf^dtv -\-\^{ dfr ^d,], _ I? frf^ f , 
00 00 00 00 oo 



The equation of blade flapping has the form 



157 



so that I dtr is proportional to the sum — -§— + p - const. Therefore, 
J di|f 



2it 1 



+jP«fp-constJrfp=i(^y| +± H _constp| =0. 

'o 'o 

dB 
since p and ,P at ij; = and \|r = 2n have an identical value [V.E.Baskin gives 

such a derivation of eq.(3.63)]« 

Thus, the expression for % can be represented as 

27t 1 _ 27t 1 

m^^llc^/J^bdrdif -^] d'ifl dti-t^V. (3-64) 



For sinplic.ity, we derived the expression for m^ on the basis of eqs.(3«50) 
and (3.51) for U^ and Uy . More accurate expressions, taking account of the com- 
ponent of induced velocity v^ , 

t7^=7 + (Fcosa-'Uft)sin^; (3.65) 

Z7, = Fslna-z)<-(i7cosa — t)Jsinpcos<l<-r ^, (3.66) 

will yield, after analogous conputations, the following expression for m^: 

/w^ = J J c.^bO^ d7d^ + S^d^l {dt^, - dt^,) -t,V. (3.67) 

00 00 

The first integral in eq.(3.67) contains the forces of profile drag and the /1^9 
second the forces of induced drag. We designate them, respectively, as 

2it 1 





2it 1 



For an ideal rotor for which c^p = and for which the induced velocities 
are loniformly distributed, we obtain from eq.(3.67) with an approximate consid- 
eration of tip losses, 

158 



or, multiplying both sides of the equality by the loading factor, 

mt=^[CyVy~C^(V-v,)]. (3.68) 

Equation (3.68) coincides with eq.(3«14) obtained in the ideal rotor 
theory. 

We note that in the expression for m^ used for calculations in the classical 
theory, the term taking account of the profile drag of the blades does not 
coincide with that obtained above and, in conformity with eq.(3.55) and (3.57)» 
is equal to 



2it 1 



m-p, 



pr = f f CjcpU U / b dr d^. 







The discrepancy of these expressions is due to the fact that the profile 
drag forces enter not only into the expression for _ ■- but also into the ex- 
pression for — ^ and thus into t^ and ty as well; if c^p is taken into account 
dr 

at some fixed values of the angle of attack a and the pitch cp of the rotor (i.e., 
true angle of attack of the section as is done in calculations by the classical 
theory), then both m,. (by an amount mp^ ) and t^ and ty also will change. When 
calculating mt by eq.(3.67), the term mp^ determines the increment in m^ , pro- 
vided identical values of t^ and ty are maintained, which obviously occurs at 
different o- and cp for rotors with different c^p . 

Since it is of greater interest to conpare rotors with different profile 
drag of the blades at identical t^ and ty, the profile losses of the rotor are 
estimated with respect to the quantity mp^ calculated at angles of attack and 
pitches of the rotors corresponding to the same value of the coefficients t^ 
and ty . 

For the reasons presented a.bove it is obvious that, for changing from an 
ideal rotor with certain ty and t^ to a real rotor with the same ty and t^ , the 
profile losses must be determined from the expression for mp^ ♦ 

It also follows from eq.(3.67) that the influence of a nonuniform induced 
velocity distribution over the rotor disk at given ty and t^ is directly deter- 
mined by the quantity m^^^ . Fiirthermore, the form of the induced velocity dis- 
tribution influences the angles of attack of the blade sections and thus also 
the quantity mp, . 

The flapping angle of the blades does not directly enter into the co- /150 
efficient m^ owing to the fact that the integral (3.63) is equal to zero. How- 
ever, flapping does influence the distribution of c^p , dt, and v over the rotor 
disk and hence the quantity mp^ and mi„4 . 



159 



3. Rotor Profile Losses 

As shown above, at given values of the coefficients ty and t^ the rotor 
profile losses are deteraiined 1^ the expression 

2x 1 



'"^"^n ^'"^^^ ^'' ^^- (3.69) 







With consideration of the radial velocity conponent of flow past the blade 
(see Fig. 2.91), the profile losses are determined by the e^^jression 



2it I 



m^r = f J c^p (U^ + K2 cos2 a cos^ <^fi^ bdrd^i. ( 3 . 70 ) 







Equation (3 '70) should be used at small values of Mq and ty , i.e., in 
cases when the coefficient c^p is determined mainly by friction forces. At 
large Mq and ty , when regions of flow separation and an increase in wave drag 
appear on the rotor, the profile drag of the sections is determined by the ve- 
locity conponent of the stream U normal to the blade axis, and eq.(3.69) must 
be used for calculating mp, . 

To calculate mp^ it is necessary to know the distribution of the profile 
drag coefficient c^p and velocity of flow U about the blade sections over the 
rotor disk. Consequ.ently, a calculation of mp^ is a laborious task and, in 
practice, can be performed only on high-speed conputers. 

Figures 2.63 - 2.70 give graphs of mp, as a function of the coefficients 

■ty> tx, V, Mq = —^ — for a rotor with rectangular blades and a loading factor 

of a = 0.091. The rotor blades (variant II in Table 2.10) have a linear geo- 
metric twist Acp = 7°» The blade profile is as follows: at the shank, up to r = 
= 0.85 - NACA 230 with a relative thickness 'c = 12^, at the end of the velocity 
profile with a relative thickness "c = 9^« The coefficient Cxp obtained on ex- 
posing the model to an airstream increased by Ac^p = 0.002 and pertains to a 
rotor having a high profile drag owing to poor manufacture. The aerodynamic 
characteristics of the profiles are given in Section 4, 3« 

The calculation was carried out by means of eq.(3'69)- Here, it was as- 
sijmed that the induced velocity is constant over the rotor disk and was deter- 
mined by eq.(3«46). 

The method of calculation and the remaining assunptions are described in 
Section 4 '2 and 4«4» 

We see from the graphs that at numbers Mq =0.6-0.7 the quantity mp^ 
greatly depends on ty and t^ • 

At Mq ^ 0.5, the quantity mpp depends little on tjj and increases somewhat 
upon an increase of ty . 

160 



nM 




Fig.2.63_ Coefficient of Profile Power of Rotor 
(V = 0.15; Mo = 0.6; a = O.O9I). 



t^ 














1 






\ 
















\ 
















' '- 


^ — 




I 


■~ 




\. 


















\ 










\, 


















\ 








^ 




_ 


f 


, 


^ 


\ 








\ 
























>" 


|. - 


\ 








\ 






















\, 




















\ 




1 












S 
























0.011Z 




1 


om 


\ 




\ 


0.0 


01 






\ 


0.005 








O.OOB 


m.pr 






ty-O 


OS^]. 












\ 










90.-5' 
A fl* 
. -J* 

A -10' 
X -15' 
















\ 




\ 












V 


s?- 


















, 


\ 






\ 












\ 


h 


3 


— 


/7/77 










^ 


^ 






> 


,■=; 












\ 










\ w 




v? 






\ 














\ 












Ir 




\? 


=1 






^ 














\ 




















^_ 














\ 1 





Fig. 2. 64 _Coefficient of Profile Power of Rotor 
(V = 0.2; Mo = 0.6; a = O.O9I). 



161 



t> 




A 






—- 


— 


— 




— 






1 










' 












V ' 






1 






' 




-- 




1 




1 

QOOJi 

- 


a 


, 












' 








lycw 














t,-c.ot 


1 
















\ t^O.Oi 


'\ 


















\ 
















I '^ 


oov\ 








O.0OS 


"V/- 




\ 




\ 




— 


1 • 


1 




, \ 




*, 


1 Ol-O 

• -/♦ 
« -«• 
« -«• 




\ 




\ 


V- 




- 








- \ 






\ 








\ 






\ 














\ 










/ 


\ ' 








\ 


\ 


- 




tAo 


'n 


\ 




— 




1/ 

-ty=^.l 

1 


rt^V 


\ 






1 


\ 












1. 



Zi^ 



(T^/ 



-OjOl 



-0.02 





A 


^ 




A0£- 


■0 


— 


••o 

-1 










• -5' 

A -to' 
« -«• 

tr -20' 










- t,-o. 


?»- 










\ 
















\ 










T 








\ 










aooull 




<?W5 




\ 




0,000 m 


\ 










V 








\ 


\ \ 










Vx 


\ 


\ 


V 










^^1 


'\ 


\ 


\ 












\ 


J 


^\ 


N 


V, 














\\ 




\ 












(f 




\ 












\ 


h 


If 








\ 


\^ 




\ 







Fig. 2. 65 Coefficient of Profile Power Fig. 2. 66 Coefficient of Profile Power 
of Rotor (V = 0.3; Mq = 0.6; a = 0.091). of Rotor (V = 0.4; Mq = 0.6; o = 0.091). 



tx 












~^ 














r 






















~' 


r 




^ 


■> _ 

\ 




























\ 
























s 


\ 






























\ 


L. 




















































\ 
































■ 
















\ 


s 
























\ 






1 


















\ 
































\ 






\ 






y 














\, 






0.002 




0.003 




\ 




0.001^ 








0.005 




\ 


K 


0.0 


06 


0.007 


mpr 
















' 






1 
















\ 












1 1 
























\ 
















\ 


^, 




ort-5* 




ty=O.0B' 


4 














\ 


c^ 












\ 


\>. 












A 

• -r 

* -w 










I 


•*1 


r*- 


— 


\ 


t 






\ 


\ 
















NT 


















\ 


«h 








V 


^ 


















\ 










X -15' 










\ 


^\-^ 


r 








\ 




















\ 








P -20' 










\ \ 




\ 










\ 
















J 




N 






L 


[_ 


J 



Fig. 2. 67 _Coefficient of Profile Power of Rotor 
(V = 0.15; Mo = 0.7; cr = 0.091). 



162 



/l^ 



ix 
0JJ2 



-0JI2 









I 


^ 


\ 






\l 1 1 










4< 






■ 




\ 


-■ 


\ 








% 




^ 








\^ 








\ 




■\ 1 i 








^1 


nP^\ 




\ejm 




6.025 




o.oot 




mpr 




1 






\ 




\ 






ort-^ 


-° 


\~^r 


'0.06 


\ 




\\ 1 




• -^ 


f 




\ 






\ 




* - 


0" 


yi I rt 




\ 


1* 




X 


% 


- - 

'r 


i' 


l>^^'* 

^t.!! 


1 


: _t^r 










V. 


^\• 


X7 




\* 






s 




?\ 1 




i\ 








\ 





Fig. 2. 68 _Coefficient of Profile Power of Rotor 
(V = 0.2; Mo = 0.7; CT = 0.091). 



U.Ui 






t 










\ 






0.0014 ' 


t 


\ 






. \ 


\ 






\ ^ 


v 








\ 






\ 


^ 






, \ 


V 






1 \ 


\^ 






\ \ 








\ o.dor^ 


0,006 \ 


mpr 




\ x*^ 




\ 










r T^ 


• -5 








\ \^^ 




'OJJi 


X X 


X -/J 

-2t 


r 




1 w ^ 


A -to 


o 




i;3' 


\ 






t^O.OB *- 
1 1 


r i^ 


^ 

k 








\ 








^.. 





Fig. 2. 69 _Coefficient of Profile Power of Rotor 
(V = 0.3; Mo = 0.7; o- = O.09I). 



163 



0.01 



-0.01 



-0.02 







A 

--- 


\~ 


- 


— 


-- 


\ 


\ 


\ 


















— 


1 




1 






r 
















\ 




0.007\ t\} 


0.008 






0.009 








^^.oto 


mpr 




' 




















\ 


fl' 1 




* 








V^ 












\ 


■i=.t 




k 


' \ 






\ 


\\ 














^ 


— 




3 








\-i 






• -5° - 

» -w — 

» -!5' 
-20° ~ 
o -25'- 

1 1 1 


? 
' 








s 


\ 








A 




\> 






ifd.06\ 


3 








s. 








\f 








\ 


\, 




\ 


i 


1 










\ 








J 



The materials for determining 
Mpj. for blades with other geometric 
characteristics are presented below 
in Subsections 4-6. 



/154 



4« Certain Considerations in Selec ting 
Blade Shape and Profile 

The power expended to overcome 
profile losses of the rotor consti- 
tutes a large portion of the total re- 
quired power of the helicopter. As 
shown in Fig .3. 52, about 50% of the 
required power is e:!q3ended for over- 
coming the profile drag of blades in 
horizontal fUght. 



Fig. 2. 70 Coefficient of Profile 
Power of Rotor (V = 0.4; Mq = 
= 0.7; cj = 0.091). 



Since, the induced losses consti- 
tute a smaller portion of the losses 
than the profile losses, we can con- 
sider that the conclusions as to the 
effect of the geometric characteristics 
of the blade on profile losses pertain also to the total power of the rotor, 
especially at high flying speeds when the induced velocities are small and the 
induced losses do not exceed 12 - 15% of the total power. 

Figures 2.71 - 2.74 contain conparative graphs of the profile drag coeffi- 
cient mpr for rotors with five blade variants: 

Variant I - trapezoidal twisted blade with high-speed profile at tip; 
Variant II - rectangular twisted blade with high-speed profile at tip 
(rotor described above); 

Variant III - rectangular twisted blade with NACA 23012 profile; 
Variant IV - rectangular twisted blade with symmetric NACA 0012 profile; 
Variant V - trapezoidal flat blade with high-speed profile at tip . 

A detailed description of all blade variants is given in Table 2.10. 

A conparison of the blades is carried out for average and large lift coef- 
ficients, at two values of V: 0.2 and 0.4. 

A conparison shows that at low Mach number M^ :S 0.5 the trapezoidal /I55 
twisted blade, at all values of the propulsive force, has approximately &% less 
profile power losses than the rectangular twisted blade. Since, in horizontal 
flight, about one half of the required power is expended to overcome the profile 
drag of the blades, a decrease in mp^ by 8% will lead to a decrease in the re- 
quired power coefficient m^^ ^ by 4%« 

Therefore, for light helicopters for which Mq is small and \ax " 0.3, the 
optimxmi planform of the blade is trapezoidal. 

The plane blade in an autorotation regime does not differ from a twisted 



164 



-0.01 



-0.02 



-0.03 



0.005 mpr 




Fig. 2. 71 Coefficient of Profile 
Power of Rotors with Blades of 
Different Shape (V = 0.2; Mq = 
= 0.6; a = 0.091). 



blade, but becomes appreciably worse 
than the twisted blade in helicopter 
regimes, especially at large V. It can 
be used for a helicopter only at small 
and average values of ty and Mq < 0.6. 

•When varying the blade profile at 
small peripheral velocities, mp, will 
vary within 5 - 1.2%. The symmetric 
profile is somewhat better than the 
asymmetric; at V = 0.2, the blade with 
a thin high-speed profile on the tip 
has smaller losses. 

We should mention that the influ- 
ence of the quality of manufacture of 
the profile on mp^ can prove to be 
greater than the effect of the type of 
profile: Amp^ for different profiles 
is about 0.0002 (the maximum difference 
at large ty is not more than 0.0004), 
whereas owing to difference in the type 
of construction and quality of manu- 
facture of the blade the profile drag 
coefficient of blade sections may dif- 
fer by an appreciable amount going /157 
as high as 0.003 - 0.004 (see Sect .4, 
3), which gives a difference in pro- 
file losses - in conformity with 
eq.(3.71) - of 



Awz, 



0.003 



/"•- 



(1+3-0.32)=^ 0.001. 



At high Mach number Mq (Mq = 0.7; tuR = 230 - 238 m/sec), the use of a high- 
speed profile at the blade tip markedly reduces profile losses . 

The decrease in mp^ amounts to 0.0015 at V = 0.2 and 0.004 at V = 0.4. 
This reduces mp^ by 40 - k5% and the total required power, by 20 and 25^ re- 
spectively. 

In flying regimes (t^ < O) of helicopters, mpp is greatly affected by the 
geometric twist of the blade. A straight blade is not used in helicopter 
regimes, and in autorotation regimes its profile drag does not differ from that 
of a twisted blade. 

The trapezoidal blade is better than the rectangular one in autorotation 
regimes and at low propulsive force. In helicopter regimes where the angles of 
attack of the tip sections of the trapezoidal blade are larger, the difference 
in mpi. decreases while at large t^ the rectangular blade becomes better. 

At moderate Mach n-umber Mq (Mq = 0.6; cuR = 197 - 204 m/sec) the peculiari- 

165 



Zi^ 



t. 












1 




\ 


M 












\ 


1 

\ 

1 




J 


\ 


\ 




















I 












'i 


2?\^ 


- 


i 


y 


1^1 














\_ 






M 


w 








aoi 












L t 




















N 


I 

1 


\ 

\ 


- 






\ 


\ 














_^ 














k 






l\ 












\ 
















\ 






yi\ 












\ 


\ 












\ 


















\ 
























It 


- 










*» 


















\ 




' 


\I 


\ 












4 




(WOz 






0.00J , 


(2eo« \ 












o.ooe 


\ 


\ 

1 


0.007 


- 




mpr 




— r 
















V 






\ 














- 




- 












\ 








\ 








y 












1 


\ 












\ 










\ 




i 


\ 












\ 


\ 








-0S1 










\ 












V I 


\. 




\ 












\' 






- 










n 


, 










\ 


\ \ 




\ 












\ 


K 
















v 










s 


1 ^ > 






\ 










\ 


\ 






















\, 












%\ 


\ 


- 


1 


1-. 
\ 










^ 


\^ 


- 




— 














nl 












H 


> 


-om 


'• -5 

' A -10 

X -15 




t - 


iy=0J6 

.ty=P.2 


\ 




V 
\ 






\ 










• 


T 


\ 




s^ 








\ 








\ 






- 




• 


jj 


\ 




s 


.\ 






\ 








\ 







Yv 














\ 


1 


s 






N 


\ 


- 




\ 


\ 


- 














^ 














\ 


V 




\ 






-BOX 












S 














^ 


J 






^ 








K. 






>3v 



Fig. 2. 72 Coefficient of Profile Power of Rotors with Blades 
of Different Shape (? = 0.2;Mo=0.7;<J= 0.091). 




-OSII 



-<uoz 



Fig. 2. 73 Coefficient of Profile Power of Rotors with Blades 
of Different Shape (V = O.4; Mo = 0.6; o = O.O9I). 



166 



ties of the curves of mp, , which were noted at Mq =0.7 (to a larger extent, at 
V = 0.1+) t begin to appear: The twisted blade with a high-speed profile at the 
tip hecomes better, and the straight blade in helicopter regimes becomes ap- 
preciably worse than the twisted blade. 

A conparison of rectangular and trapezoidal blades for Mq = 0.6 - 0.7 shows 
that at V = 0.4 and also at V = 0.2_for large ty the former has the advantage, 
whereas for medium and small ty at V = 0.2 the trapezoidal blade becomes of ad- 
vantage. In general, the rectang-ular blade is preferable for heavy and mediimi 
helicopters, whereas it is preferable to use trapezoidal blades for rotocraft 
for which the coefficients ty and t^ of the rotor are small at large V, owing to 
the installation of a wing and a tractor propeller. 




-OM^ 



Fig. 2. 74 Coefficient of Profile Power of Rotors with Blades 
of Different Shape (V = 0.4; Mq = 0.7; o = 0.091). 



At V = 0.4, the profile losses are quite large even with a high-speed /158 

profile at the blade tip: mjjj. is twice that at Mq = 0.4 - 0.5« To estimate the 
possibility of decreasing the quantity mp^ , I'igs.2.75 and 2.76 give graphs of 

for blades of the variants I and II and also for a rectangular blade with an 



nipr 

increased geometric twist (variant VI), for an expansible blade (T] = 0.5; vari- 
ant VII), and for a rectangular blade with an increase to r = 0.75 of the part 
with a high-speed profile (variant VIII). We see from Figs. 2. 75 and 2.76 that, 
in horizontal flight and especially in autorotation regime, the trapezoidal 
blade remains preferable. At large values of ty , the optimum blade is the blade 
with increased twist, which reduces mpj. at V = 0.2 by 20^ '■"^•th.f ^^ -'-'-'^) ^^ ^* 
V = 0.4 by 10^ (mt by 5%). Consequently, the use of a blade with greater 
twist raises the dynamic as well as the static ceiling of the helicopter, in- 
creases the static thrust (see RLg. 2-171), negligibly increases the maxim-um 



167 



H 

On 




-0.0t 




-m 



Fig. 2. 75 Coefficient of Profile Power of 
Rotors with Different Blade Shapes 
(? = 0.2; Mo = 0.7; CT = 0.091). 



Fig. 2. 76 Coefficient of Profile Power of 
■ Rotors with Different Blade Shapes 
(V = 0.4; Mo = 0.7; a = 0.091). 



speed, and appreciably increases the rate of descent in autorotation. An en- 
largement of the blade portion with the high-speed profile slightly reduces the 
value of nipr . 

The obstacle in using blades with greater geometric twist lies in the in- 
crease of dynamic stresses in the blade spar, whereas for blades with an en- 
larged high-speed profile, the increase of hinge moments is the obstacle. 

The expansible blade is preferable over the rectangular design only at very 
large values of the propulsive force coefficient tx • 

A comparison of the graphs of mpr for Mo = 0.6 and Mq = 0.7 shows that, in 
a rotor with our high-speed profile at the tip, it is impossible to avoid a pro- 
nounced increase in profile losses at Mo = 0.7, for all blade variants. 

The method of utilizing graphs of mpr for rotors with these types of blades 
but with a different loading factor is described in Section 6. 



5. Approximate Determination of Rotor Profile Losses 



The quantity mpr is most reliably determined from graphs plotted for each 
specific rotor. If there are no such calculations, the data of Figs. 2. 63 - 2.74 
can be used for an approximate estimate of mpr. 

At small Mo, the approximate equation (3.72), derived on the assiimption of 

constancy of the coefficient cxp in all blade sec- 
tions, can be used for determining mpr. 

Let us derive eq.(3.72). For a rectangular 
blade, we have b = const = 1.0. Having taken 





TABLE 


2.5 




1 


1 


2 


3 


4 


p 


1.0 


0.94 


0.91 


0.88 



U^ = {Ul + UlY'^^{r+Vsmif)\ 



we obtain 



m 



U • U 



Wdr = 



(3.71) 



where cxp,^ is the average value of the coefficient Cxp over the rotor disk. 

For trapezoidal blades, mpr is smaller than for rectangular blades. This 
is taken into accoimt by the coefficient P which is pre-assigned in relation to 
the blade taper T) (Fig. 2.77) in Table 2.5. 

To account for the influence of the radial velocity component of flow past 
the blade, mpr is calculated by eq. (3.70). An approximate estimate (Ref.25, 36) 
shows that, to account for this component, the coefficient of V^ in eq. (3.7I) /I6O 
should be changed from 3 to 5. 

169 



Axis of rotor rotation 



ih 



V//M/d 




n-i 



"tip 



tip 



Fig. 2. 77 For Determining Blade 
Taper. 



Thus, the final formula for detei^ 
mining mpr reads 

^p^=\^.p.^^■^^^')p■ (.3.12) 

The average profile drag coeffi- 
cient Cxpav over the disk is determined 
as a function of the average lift coef- 
ficient CyQ over the disk; the latter 

is found from eq. (3«74) whose deriva- 
tion is given below: 

2ic R 





1-K 1 

= -^ QCy, bR {y>Rf -1 J rf^ J t72 cfr. 



Substituting 



t/2;^(r + Vsin<l;)2 



and integrating, we obtain 






(3.73) 



Expressing Y in terms of the dimensionless coefficient Cy or ty, we find 

Cy 3 3 ^y 



"■yo- 



<s 3 _ 3 - 

1-1--— y2 l-f V2 

2 2 



(3.74) 



Having determined Cy^, the profile polar in the section r = 0.7 will 
yield CxPay. 

L.S.Vil^dgrube proposed to take into account the planform of the blade by 
the coefficient P and to determine Cxp ^y ^-s a function of Cy^ . 

6. Effect of Air Compressibility on Rotor Profile Losses 

At average and large Mo (for profiles generally used at Mq > 0.55 - 0.6, 
i.e., at cuR > I85 - 200 m/sec) it is necessary to supplement mpr, calculated 
from eq. (3.72), by the term Amoo which takes into account the increment in 



170 



profile power produced by the increase in profile drag coefficients of sections 
over which the flow has high Mach numbers. Thus, /I6l 



m^, = 4-r,„ (l+5l/2)P+Am, 



(3.75) 



W; 



The coefficient of the increment in profile power Amco shoiild be determined 
with consideration of the actual distribution of the angles of attack of the 

blade sections over the rotor disk, 
since an increment in profile drag due 
to an increase in Mo generally occurs 
in all blade sections. Figures 2.78 
and 2.79 show M.M.Tishchenko»s graphs 
of the variation in the profile power 
coefficient of a blade as a function 
of its azimuthal position in the 
plane of rotation. We see from 
Fig. 2. 78 that, at low flying speed 
(V = 0.2) but at large thrust coeffi- 
cient, the profile power of the blades 
increases at all azimuths as the Mq 
increases. At high flying speeds 
(see Fig. 2. 79), the increment in pro- 
file power occurs mainly at azimuths 
of 30 - 150°. 



OfitS 
0.01 



oms 













■ 1 

MifOl 








V-'0.2;t/-0.Z 








/ 


- 


< 




0.65 
















/' 


^"^ 
















X 






^- 


K 


>< 


y ■ 


30 
















^ 


^ 






■^ 


^ 








-<J 


^ 


— 


— 




^ 








-— 


' — 


Z::! 





















wo 



ZOO 



300 



d)' 



Fig, 2. 78 Variation in Profile Power 
Coefficient of Blade with Respect to 
Azimuth. 



The graphs of Amoo for the variant II of the rotor are given in Figs. 2. 80 
to 2.84. The quantity Amoo is defined as the difference between the profile 
power coefficient_mpr at the examined Mo and at Mo = 0. 4 at identical values of 
the coefficients V, ty, tx : 



A/n,„ (Mo) = m^r (Mo) — /«a^ (Mo = 0.4). 



(3.76) 



It follows from Figs. 2. 80 - 2.84 that Amoo is a fimction not only of V and 
the Mach niomber Mq but also of the coefficients t-^ and tx . The coefficients ty 
and tx have an especially strong effect at small V at which, in conformity 
with Fig. 2. 78, the increment in mpr occurs at all azimuths. Upon an increase 
in V the increment in mpr occurs mainly in the region 'I' = 90° (see Fig. 2. 79) 
where the angles of attack of the sections are close to zero regardless of the 
value of ty. Consequently, at V = O.4 and V = O.5 the influence of ty and tx 
on the quantity Amoo is insignificant. 

We see from Figs. 2.80 - 2.84 that, at large Mo, V, and ty, Amco is large. 
The quantity Amoo ^eatly increases when Mo > 0.55 - 0.6. At near-separation 
values of ty when V = O.I5 and V = 0.2, Amco has a high value already at 
Mo > 0.5. 

So as to keep the increase in required power of a helicopter, due to the 
compressibility effect, from exceeding 15 - 18^, the rotor of the variant II 
should be used when Mo =_0.7 at V ^ O.3, and when Mo = O.65 at V ^ O.4. For 
example, when Mq = 0.7, V = O.3, ty = O.I5, and tx^^^ = -0.0075» the increment 



171 



/162 



0.035 
0.030 
0.025 
0.020 
0.015 
0.0) 
0.005 

















V'-O.l; by"0.11 










\ 






























\ 


























\ 




\ 


























\ 




\ 


























\ 






























1 


r 


\ 
























1 




\ 


L 


.^M 0=^0.7 




- 


- 






1 




\ 


\ 


_^ 


^ 


0.65 














\\ 




^ 


0.6 






Jj 


/ 




'\ 


j^ 


■^ 


^0.55 














// 


'/ 




N 


^ 


\ 














^ 


y 








"^ 


)\ 












.-- 






f 












^ 


U 




^ 


^ 

i' 


j^ 
























--' 









100 



200 



300 (6' 



Fig. 2. 79 Azimuthal Variation in Profile Power 
Coefficient of Blade. 



V'O.IS 




)0.W 

]troiB 



Fig. 2.80 Increment in Profile Power Coefficient of 
Rotor, due to Air Conpressibility. 



172 



Z162 



0.001) 




1$>0,16 
0,1 Mo 



Fig. 2.81 Increment in Profile Power Coefficient of 
Rotor, due to Air Compressibility. 



173 



^064 




op Mo 



Fig. 2.82 Increment in Profile Power Coefficient of 
Rotor, due to Air Compressibility. 



174 



/165 



"CO 



0.005 
0.001 
0.003 
0.002 
0.001 



V =0A. G =0,091 

























J 






" r r I T II 1 " 


- 








h- 




-- 


- 








- 


-4 


1 


\ - 




.; 








i 


I 


— f ~ 


h 




r_L.. 








A^ 


m^- 






-j 


- 


. . 




g 


^ 


^ 


i 


'^ 


— 



3\ty-0.tS 

\0.1U 
0,10 




Fig. 2.83 Increment in Profile Power Coefficient of 
Rotor, due to Air Compressibility. 



175 



^'"co 








y^Q3; 6='0,091 








n 

/ 


009 










































om 
























/J 
n 
























0.007 






















1 

1 


_ 






















0.00S 










































0.005 




















) 


f_ 























OfiO'i 




















1 


— 




















0.003 




















1 


- 


-- 


















/ 




0.002 


















/ 
















/ 


0.001 


































y 






— 


— 












^ 


^ 




_ 







^66 



at/ 



H5 



0.S 



VTM, 



Fig. 2.84 Increment in Profile Power Coefficient of 
Rotor, due to Mr Compressibility. 



4a„ 



-10 ■ 



-2.0 - 



-3.0 . 



0.1 




0.5 






0.6 






0.7 Mc 


i 









- 


-~ 


^^ 


~T 


. 


^ 


ty=0.lS\ 


1 






^ 


<-/- 


u.iv 


j 






'"N 








: I 




7=«<l; 6=0091 












- 






ty^O.W \ 


1 






- 




- 


- 


- 


i 











Fig. 2.85 Increment in Angle of Attack of Rotor, due to Air 
Compressibility at Constant Coefficient of Propulsive Force. 



AOlr. 



-1.0 



-2.0 



-3.0 



0.Q 






0.5 





















"^ 




























V-O.S; 


6=0.091 














- 





























0,6 



N 



0.7 Mo 



\ 



\ 



\ 



Fig. 2.86 Increment in Angle of Attack of Rotor, due to Air 
Compressibility at Constant Coefficient of Propulsive Force. 



176 



in profile power vd-ll be AiUoo = 0.0016 which amounts to IBfo of nit^.f • When 
Mo = 0.65, V = 0.4, ty = 0.13, and txj^.f = -0.0133, the increment in profile 
power will be Amco = 0.002 which amoiints to 16^ of mtj,. f . 

Since, at large 7, the increment in mpr ;jr occurs mainly at azimuths close 
to 90 , the increase in Amco at large V is intimately connected with the rela- 
tion between the Mach niomber of the blade- 



Am 



0,001 



■0.003 




0,002 



0,001 



Fig.2.S7 Increment in Profile 
Power Coefficient, due to Air 
Con^sressibility for Rotors with 
Blades of Different Shape. 



tip section at i^ = 



90°, 



equal to Mn + Mo = 



= Mo(l + V), and the critical Mach number /167 
of the section profile. The critical tfech 
number Mcr is determined at Q? = since, in 
the tip sections at ilf = 90°, we have oTj. « 0. 



These data show that a 15 - 18% in- 
crease in required power, due to the comr- 
pressibility effect, occurs at Mfj + Mo = 
= 0.91, i«e., Mfi + Mo is larger by 0.1 
than Mcr of the high-speed profile when 
or = (see Fig. 2.99). At Mn + Mq = M^r + 
+ 0.15, the increase in required power is 
about 30^. At Mfi + Mo = Mor, the compres- 
sibility effect is Arirtually absent. These 
relations between Mq, Mfi , and Mor of the 
blade profile can be used when selecting Mq 
for a helicopter with high flying speeds. 



Since, on an increase in Mq, the angle of attack of the rotor shoTild be 
more negative so as to retain identical values of the coefficients ty and tx , 

the graphs for the increment in rotor angle 
v=o.«; 6=0.031 of attack are given in Figs. 2. 85 and 2.86: 



Am CO 
0.008 

0.007 

0.006 

0.005 

aoon 

0.003 
0,002 
0.001 



tx =-0.010 




m 




_/ 




^ 


ty = O.W 


// 




ty = 0.t1 


/ 






V 






7 






/ / 


u 




h 


y. 




ri 




h 


/// 


n 


t 


7/ 




j^ 


^- 




~^t-7^ 




^^5^ 






^^i^ 






_=^=^£ = 







0.f 



0.5 



0.6 



07 Mg 



Fig. 2.88 Increment in Profile 
Power Coefficient, due to Air 
Conyressibi li.ty for Rotors with 
Blades of Different Shape. 



Aa„ (Mo) = a(Mo)-a(Mo=0.4). 



(3.77) 



No graphs were constructed for V « 
= 0.15 - 0,3 since, at all ty and tx, the 
quantity Affoo does not exceed 1 . 

Figures 2.87 and 2.88 show comparative 
graphs of Amco for rotors with blades of the 
variants II, III, and V. Calculations 
showed that the quantity Amoo is greatly 
affected by the type of profile (this is 
seen from a con^iarison of the variants II 
and hi) and by the geometric twist of the 
blade (variant V, straight blade). The 
planform of the blade plays a role only at 
large ty in which case, for trapezoidal 
blades where flow separation begins earlier, 
Amoo is greater than for rectangular 
blades. The planform of the blade plays a 
minor role at large Mo in view of the fact 



177 



II 



that, as will be shown in Section 4«7, a variation in planform will cause a 
change in the angles of attack of the sections mainly at azimuths ^ = 250 - 340° 
where profile losses are small in the pre-separation regime. 

A comparison of blades with a high-speed profile at the tip (variant II) 
and without it (variant III) used in the calculation, will show that a high- 
speed profile must be established at the blade tip when Mo > 0.6 - 0.625* 

The graphs of Afflco and Affoo should be calculated for each specific rotor. 
However, if no such calciilations are available, the data in Figs. 2.80 - 2.88 
can be used for an approximate estimate of Amoo. 

The graphs in Figs. 2.80 - 2.88 are laid out also for taking into account /168 
the influence of the Mach ntimber Mq on the aerodynamic characteristics of a 
rotor, for cases in which the rotor characteristics experimentally determined 
at low Mo are to be used also at high Mq. Furthermore, the graphs are useful 
for aerodynamic calculations to avoid interpolation of the Mach mm±ier Mq if the 
calctilated Mo values do not coincide with those for which the graphs of the 
aerodynamic characteristics were plotted. 

To use the graphs shown in Figs. 2. 80 - 2,S3 for rotors with similar blades 
but with a different loading factor, it is necessary to recalculate the coeffi~ 
cient tx for a = 0.091 (Sect. 6). 

7. Induced Losses of a Real Rotor 

Assuming a constant induced velocity .over the entire rotor disk, the torque 
coefficient can be obtained from the graphs in Rigs.2.60 and 2.61. The inter- 
dependence of the ratios Cy/B^V^, Cx/B^V , and mt/V°, which was derived in the 
theory of an ideal rotor, is valid when these ratios are determined with respect 
to the total forces Y and X taken with consideration of the profile drag, since 
the forces of the profile drag also create induced velocities so that the ve- 
locity polygons and all relations given in Subsection 1 remain in force. We 
must add the profile losses to the mt obtained in this manner. 



Consequently, 



V 
where 



m,==~^i-V'^m,., (3.78) 



The addend in the expression for St, containing the product of the aerody- 
namic force and the induced velocity CrU, will be called the induced losses of 
the rotor. 

In calculating the induced losses, we introduce a correction for taking 
into account the nonuniform induced velocity distribution over the rotor disk. 



178 



As follows from eq. (3»67)» the induced losses of a real rotor are deter- 
mined by means of the formula 

27: 1 _ 2k I _ 

"I ind = J ^=^1* j dt^u = J rf^p j {dtyVy -dt^v). (3 . 79 ; 



However, to calculate ordinary helicopter regimes, at ty > tx, an approxi- 
mate expression is used 

m^^^j'^cl<^'^dtv. (3.80: 



First, just as in the ideal rotor theory, we determine mi„ii under the as- 
sumption of constant induced velocity over the entire rotor disk. With this as- 
sumption and with an approximate consideration of tip losses, the expression 
for mind takes the sin^jle form 



m 






For flying regimes at V ^ 0.15, • substitution of eq. (3.4S) for v will yield 

mind=^. (3.82: 

We will demonstrate that eq. (3.82) holds not only for the assumption of 
constant induced velocity over the rotor disk but also for an induced velocity 
distribution obeying the law 

v{r,'^) = v-\-arcos<\i, (3 •83, 

where a is a constant. 

According to eq. (3.83), the induced velocity has a minimum value in the 
forward portion of the rotor disk (If = tt, r = 1) and increases linearly in the 
direction of the velocity flow. In a direction perpendicular to the velocity 
flow, the induced velocity remains constant. Thus, the form of the induced ve- 
locity diagram is a cylinder cut off by a plane turned toward the plane of rota- 
tion of the rotor about an axis perpendicular to the direction of motion (see 
Fig. 2. 6); the angle of turn is characterized by the quantity a; v, the average 
induced velocity of the disk, is determined from eqs.(3«46) or (3«48)« 

The induced velocity diagram described by eq. (3«83) is close in character 
to the time-average induced velocity diagram found from experiment (see Fig. 2.3. 

Let us substitute eq. (3*83) into eq. (3.8O) and find 

2nl__ _2ill 2r. 1_ 

mjnj= Cdi]) ^ dt{v-{-ar COS <Sf)=v f rftj) frf^ + a ^cos'^id'^ l'^^''- 
06 6060 

179 



The integral in the first addend is equal to the rotor thrust coefficient. 
The integral for radius in the second addend is proportional to the sum (see 

Subsect.2) — -5— + P - const for a rotor with flapping hinges. With an accuracy 

d^B ^ 

to the first harmonics of flapping, the sum 5- + P = aoj consequently, fdtr 

is a quantity independent of the azimuthal position of the blade. Therefore, 

2TT X 

Jcos ijr d'ifjdtr = 0. 

o 

Thus, for the induced velocity distribution in accordance with eq. (3.83) 
the induced losses are also determined by eq. (3«Sl). Calculations based on the 
vortex theory for a rotor with an infinite nxmiber of blades show that, owing to 
differences in the induced velocity diagram from eq. (3.83), the induced losses 
of a rotor with twisted blades are about 5% greater. Taking B = 0.92, the in- 
duced losses of the rotor in flying regimes at 7 ^ O.I5 are determined by the 
expression 

^,^, = _L:0^ -^^0.285-^. (3.84) 

'"° 0.92-4 V V 

In flying regimes with small V, the quantity mind is found from eq. (3.85) /170 
where u is determined as a function of Cr, V, and 5 from the graph in Fig. 2.62: 



— 1-05 — 1.05^3/2 ~ n t;cr>3/C 

where, just as in eq.(3»84), I.O5 is a coefficient taking into account the in- 
crease in induced losses. 

The dependence of u on V can be refined by flight tests. After determining 
the reqiu-red rotor power from flight tests for a nianber of horizontal flying 
speeds and after calculating the parasite drag of the helicopter, we find Cy^.f , 

C^Xh.f (see Chapt.III, Sect. 1.2; as well as Cr, mt^.f , and then V and u from the 

expressions 



/ 



452 



-__;%., -'"'"•+i-c^M^ 



0.56 Cf ■ (3.87) 

The graph of u = f (v) obtained from flight tests of the Mi-4 helicopter is 
shown in Fig. 2.89. 

The tests were performed at different heights between rotor and surface of 
the airfield h. In flights close to the groiond, the quantity u was affected by 

180 



the "air cushion". The values of u in the influence domain of the "air cushion" 



are plotted on the graph for different values of h = . 

R 

The graph of u in Fig. 2.89 was obtained for low horizontal flying speeds, 
when S = &^,t « 0. However, the diagram can be used approximately for values 
of S within limits from +5° to -30° . . Calculations made from the graph pf u, 
shown in Fig. 2.62, revealed that instead of determining the product uCr ^we can 
calciilate the product uC? (Cy = Cr cos S), determining u for 6=0 and V = 



= 1.96 



V 



H''7h)^^-"t''-^) 



cT. 



(3.88) 



Thus, in calculating the rotor characteristics at low flying speeds, mina 

is determined from the expression 



10 




<: 
^ 

/ 


-f 




















































.> 


V 


























- 




1 


^ 










































0.7S 


\J 


.\ 


\ 






















^ 




\ 


\^ 


\ 




















— 






N 


s 


\ 


4 




















-- 




\ 


S 


\^ 


k 


















\ 


\ 


^ 
\ 


^ 














0.50 


N 














N 


V 














— 


\ 


V 


^ 














V 


^ 


^ 






























iSa 


1^ 


0.25 


i 




















- 






































































- 





































m 



ind 



= O.SGiiCy-. 



(3.89) 



t.o 



z.o 



3.0 V 



Fig. 2.89 _Induced Velocity u as a Function 
of V and h (Based on Flight Tests of the 
Mi-4 Helicopter). 



Fig. 3. 8) close to y^ = 0.057, just as for the Yak-24. 



Equation (3.89) can be used 
both for calculating the rotor 
power reqtiired for horizontal 
flight at low flying speeds and 
for determining the propiilsive 
force of the rotor when calcu- 
lating the takeoff distance of a 
helicopter or the towing force of 
a towing helicopter. These 
calculations are substantially 
simplified because of the fact 
that, for determining the veloci- 
ty coefficient 7, it suffices to 
know Cy and not Cr . 

The graph of the average in- 
duced velocity u for a rotor 
system (with consideration of 
mutToal interference) of the 
Yak-24 fore-and-aft helicopter is 
shown in Fig. 2. 90. Figure 2.90 
also contains the curve u for the 
Mi-4 helicopter outside the 
earth »s influence. This graph 
can be used approxima'tely for de- 
termining u of all helicopters 
of single-rotor configuration /171 
and of fore-and-aft helicopters 
■with an excess of rotors yv (see 



181 



Turning to these ciirves, we can find u for fore-and-aft helicopters with 
other y and also for helicopters of side-by-side configuration, after determining 
u at high flying speeds (V ^ 0.15) from eq.(3.87), based on data of an aerody- 
namic calculation. Such a ciu-ve is plotted in Fig. 2. 90 for a helicopter of 

side-by-side configuration with a 



10 


^^ 


. 

























- 


- 


S 


^ 






\ 


\ 


\ 












■- 






\ 


<^ 




' 


















\ 


;> 


N 


\, 




Yak-2^ helicopter 
' (fore-and-aft 
configuration) 












s 


\ 


— 


V 


< 


0J5 










\ 




\ 












S^ 


s 






N 


k 






















\ 


\ 










\ 






















N 


k 


^ 








'v 
































0.S0 














\ 


s. 


N 


K 


M£-4 hel 


icopter 
















\ 


configuration) 


















\ 


k 


























\ 


X 


















1 1 y 

Hel icopt 


:r o 


K 










--^ 


0.Z5 












side-by- side 


V 




















(X^^-O.V) 


■^ 






^ 


























































— 


— 

























































coefficient 

Hss = -O.4. 



of mutual induction 



It_is interesting to note that 
at low V for a fore-and-aft heli- 
copter the induced velocity coeffi- 
cient, owing to the mutual inter- 
ference of the rotors, is greater 
than in a hovering regime. Conse- 
quently, its required power at low 
flying speeds is greater than in 
the hovering regime. 

Thus, a fore-and-aft heli- 
copter has poor flying character- 
istics at low flying speeds (in ac- 
celeration, in takeoff runs when 
taking off Hke an airplane, and in 
towing) ; they are substantially 
worse than those of single-rotor 
and side-by-side helicopters. 



V 



Since u depends on the ratio 
:, which is directly propor- 



to 



2.0 



10 V 



tional to the ratio 



V 



^/p~ 



(p being 



Fig. 2. 90 Induced Velocity u vs. V for 
Helicopters of Various Configurations. 



the load per square meter of the 



rotor area, p 



-ip), the flyini 



speed has a different effect on the required power for helicopters with dif- /172 
ferent p. Therefore, for helicopters with a larger p, the wind in this case 
lowers the required power less or increases the maximum rotor thrust in hovering. 

-!!- * -x- 

Thus, for calculating the torque coefficient of a lift-producing rotor and 
a propulsive force with coefficients ty , t^ exposed to an air stream with di- 
mensionless velocity V s 0.15, we can use the following expression: 



'"/■ ^^ '"/"• ■ 



1.05 ty" 1 



\B^ V 



2^.^ = 



(3.90) 



=/ra^r+0.285-4^ 1.04<^V/. 



182 



Owing to the necessity of taking tip losses into account, the coefficients 

B^ are often omitted in the term t^V so that eq.(3.90) takes the following 

form: B^ 



OT^ = /n/v.+ 0.285^: 1;/. 



(3.91) 



The coefficient mp^ is determined as indicated in Subsections 3 - 6 of this 
Section. 

8. Determination of Angle of Atta ck and Pitch of Rotor /173 

The momentum theory gives no data on the angle of attack of the rotor. 
Determination of the angle of attack of the rotor a and its characteristics in a 
rotor-fixed coordinate system (forces T, H), however, is necessary for calculat- 
ing the rotor pitch, for refining the magnitude of parasite drag of the non- lift- 
producing parts of the helicopter, and mainly for determining the equilibrium 
conditions of the helicopter moment relative to the center of gravity (balancing) 
and its stability. 

It is obvious that, when forces with coefficients Cy and Cx are generated 
during some operating regime of the rotor, the determined mean dimensionless in- 
duced velocities Vy and v, over the disk must correspond to these operating con- 
ditions. However, the angle of attack of the rotor may differ here and depends 
on the type of rotor (hinged or rigid), on the blade shape, etc. To determine 
the angle of attack of the rotor use must be made of the classical theory, 
wherein the found magnitude of the angle of attack depends on the assunptions 
contained in this theory. 

Let us determine the angle of attack and pitch of the rotor. 

To each point of the ciirves of mp^ (see Pigs. 2. 63 - 2. 74) there corresponds 
a certain rotor angle of attack. The angles of attack are laid off on these 
curves so that, in determining mp^ from Figs. 2. 63 - 2.74, the angle of attack of 
the rotor can be located. The rotor setting is found from graphs of ty = f(a. 
Bo, V) (see Figs. 2. 115 and 2.116) or from eq.(2.50) of the Glauert-Lock theory 
(Sect. 3), wherein p, and X are detei^nined from eqs.(3.20), (3.2I). 

If the profile power coefficient is determined from eq.(3.72) rather than 
from the graphs in Figs. 2. 63 - 2.74, then a, which is the angle of attack of an 
equivalent rotor (see Sect .2), is calculated from the approximate equation (3*95). 
This formula is derived on the basis of the following relations: 

t^ = ts.\nag-\~heCOsae^tag^k^. (3-92) 

Assuming h, ~ tai , we find from eq.(3.92) 

««~^-av (3.93) 

183 



The flapping coefficient ai^ can be e^qjressed by the approximate relation 
[eq.(3.94)] derived from formulas of the Glauert-Lock theory: 



«'.=2^K-£-w)-H- (3.94) 



After transformation of eqs.(3.93) and (3 •94), we obtain the formula for 
determining a, : 



1 



1—21/2 



T-'Hi.-^)]- (3.95) 



At large Mach numbers Mq for V s 0.4, the increment in angle of attack is 
found from the graphs in Figs. 2. 85 and 2.S6. 

After ffo is determined, ^J,, \, cp, and other data are found. 

Having determined a, eqs.(3'l5) and (3.17) will permit finding the coeffi- 
cients of thrust and longitudinal force of the rotor. 

Section 4« Classical Rotor Theory. Method of ^174 

Numerical Integration 

When calculating the aerod3niamic characteristics of a rotor in regimes with 
large V, Mq, and ty, many of the assutiptions of the Glauert-Lock theory lead to 
substantial errors. For commonly used rotors, we can consider that V ^ 0.3-0.35; 
Mo ^ 0.55 - 0.6; and ty close to ty , based on the condition of flow separation. 

In calculating such regimes it is primarily necessary to discard the ap- 
proximation of the profile characteristics stipulated in the Glauert-Lock theory: 
Cy = acoQ?r and c^p = c^p , where am and Cxp are constants at all points of the 

disk regardless of the angle of attack a^ and the Mach number of the blade sec- 
tion. 

In practice it is inpossible to give a sufficiently accurate analjrtical 
expression for the dependence of Cy and c^p on a and M. Therefore, in the re- 
fined calculation methods the angle of attack and the Mach number are found at 
each point of the swept disk after which Cy and c^p are determined from the 
graphs of the profile characteristics. 

For calculating the distribution of the angles of attack, the flapping 
angle of the blade p must be known; however, this can be determined only if the 
thrust moment relative to the flapping hinge is known. The latter can be found 
when the distribution of the angles of attack is known. Therefore, the calcula- 
tion can be constructed either on the basis of determining, by the method of 
successive approximations, the flapping coefficients with respect to the first 

2-3 harmonics, or on the basis of determining p and -4|- by numerical integra- 
184 



tion of the equation of flapping; the second method of calculation, which has 
become widespread, will be described below. 

Practical application of such a laborious conputational process is possible 
only with the use of high-speed digital computers. Under this condition, the 
previously used assimptions for overcoming mathematical difficulties can be dis- 
carded. Unavoidable assunptions are only those due to our lack of knowledge of 
individual problems at the present state of art of rotor aerodynamics. Such as- 
simptlons include: 

Determination of Cy and Cxp of sections, neglecting the angles of side- 
slip (equal to — =— j and variations in the boundary layer produced by 

centrifugal forces arising on blade rotation; Cy and c^p of the sections 
are determined from the aerodynamic characteristics of the profile ob- 
tained in a plane-parallel flow. 

Neglect of the effect of unsteady circulation of flow about the blade 
sections, which involves a complex motion, on the aerodsmamic character- 
istics of the profile. 

Neglect of the fuselage and hub effects on rotor aerodynamics, and 
others . 

The method of calculation permits taking into account (within the assunp- 
tions given above) individual features of the blade profiles and to select a 

profile on the basis of quantitative 
data rather than of qualitative con- 
siderations, as was done previously. 

The aerodynamic characteristics 
can be calculated together with calcu- 
lation of blade deformation and with 
consideration of the induced velocity- 
distribution caused by a vortex system 
of arbitrary form; the conputational 
effort depends on the accuracy require- 
ments and on the capability of the 
conputer such as memory capacity and 
speed of confutation. 




VCOSoC 



(r-lhj,)oospH^ 



Vcos^cosfp 



1. Formulas for Calculating Forces /17^ 
and Moments_ of a Rotor 



Fig. 2. 91 For Detennining the Comr- 
ponent of Relative Velocity of 
Flow around the Blade Section. 



First, let us derive formulas for 
determining the conponents of the rela^ 
tive velocity of flow about the blade 
sections. These differ from the 
formulas derived in Section 2 in that they take into account the spacing of the 
flapping hinges and do not consider the angles P and $ to be small. 

The conponent U^ (Fig. 2. 91a, b) is directed perpendicular to the blade axis 
and is located in a plane parallel to the plane of rotation (or located in the 



185 



plane of rotation when the flapping angle of the blade relative to the flapping 
hinge is zero). As shown in Pig. 2. 91, U^ is conposed of the projection of the 
fljd-ng speed, equal to V cos a sin i)!, and the peripheral speed of the section 
tuC^ - K.%) cos p + Ih.h 3: 



Changing to relative qualities, we obtain 



(4.1) 



where 



t/,=^=(r-/^^,)cosp+/^^, + \/cosasin<|.= 



= (^ - ^h.h) <^os p + /^ ,, + IJ. sin <)), 



V COS a Tr 

fi = - — =Kcosa. 



(4.2) 



(4.3) 



strictly speaking, with consideration of flow stagnation in the region of 
the rotor equal to the induced velocity y^, the flow velocity in the plane of 
rotation is equal to V cos a - Vjj . Therefore, the dimensionless coefficient p, 

must be determined from eq.(3.20). This in- 
troduces no conplications if the calcula- /176 
tion is made at a given |j,, and the dimen- 
sionless flying speed V is determined from 
eq.(3.20) when Vjj is already known. If the 
calculation is made at a given V, then for 
sinplification we will determine (j by the 
approximate equation (4.3). 




Longitudinal axis 



Fig. 2.92 For Detennining the 
Position of Blade, Hub, and 
longitudinal Axis of the Heli- 
copter Relative to the FUght 
Direction. 



and through an angle i]/ + Cq - Ps 



In this Chapter, we will not discuss 
blade flapping relative to the drag hinge. 
The variable part of the angle of deflection 
of the blade about the drag hinge is negli- 
gible and it can be considered that all 
blades turn about the drag hinge through an 
identical angle §av = Co* Therefore, at 
some azimuthal position of the blade i|;, the 
rotor hub is turned through an angle ijf + Cq 
toward the projection of the fljd-ng speed 
to the longitudinal axis of the helicopter. 



if the craft is flying with sideslip (Fig. 2.92). 

The conponent U^ (see Fig. 2. 91a., c) directed along the blade axis, is 
equal to 



U^^^V cos a cos i)) cos p. 



(4.4) 



This conponent determines the angle of sideslip in flow through the blades. 



186 



The conponent Uy is directed perpendicular to the blade axis and is located 
in the blade flapping plane (being parallel to the shaft axis when the flapping 
angle of the blade is equal to zero J . 

As shown in Fig. 2. 91c, the conponent Uy is conposed of the following speeds: 
projection of the speeds perpendicular to the plane of rotation of the 
rotor, V sin of - Vt ; 
projection of the conponent of flying speed, V cos oi cos ilr ; 

peripheral speed of flapping, (r - tj^.^) "^P -. 
The sum of these speeds is equal to 

L'^y = ( 1/ sin a — ■«) cos S — 1/ cos a cos ij) sin p — (r — / j,},) — , ( A- . 5 ) 

where v is the induced velocity conponent perpendicular to the plane of rota- 
tion (vt in Sect .3) • 

On replacing the differentiation with respect to time by a differentiation 
with respect to the angle of blade rotation (i|; = uat) and changing over to rela- 
tive quantities, we obtain 

— i/y _ _ _ 

L^y = — — =(V^sina — i;)cos p — V cosacos^sin p — 

(4.6) 
-('■-W^ = M'-,'l')cosp-i.costl;sinp-(r-/;,.^)^. 

Here the flow coefficient X(r, i; ) at nonuniform induced velocity distribution /177 
is equal to 

X (r, ip) == V sin a — 1) (r, 'jj) = V^sin a — ['u — A'y(r, (]>)] = 

(4«7) 

=.X-|-A^(r,<l-), 
whence 



>.= V/'sina — v, 
where 



(4.7') 



V and Av (r, ilf ) = mean and variable portions of the dimensionless induced 
velocity; 
\ = average flow coefficient over the disk. 

The geometric sum of the conponents U^ and Uy is equal to the relative flow 
velocity through the blade section in a plane noiroal to the blade axis: 



u=\^ui+ni. (4.8) 

The angle P and the angular velocity ■■ P .* of the flapping motion of the 

187 



blade, which are detennined from the flapping equation, enter the ejqjressions 
for U,, Uy, U,. 

Without the sinplifjo-ng assunptions made in Section 2, the flapping equa- 
tion has the form 

rf2p 



^H.h '"^ ^+(^Kh ^os P-'h.1,\h>^sln p = 






'h.J, 



or, in dimensionless fonn. 



where 



rf.l/2 ~V ^ hh ' «- ^•'* ^vv<«2 



'Kh 



0)2 



QbojR* 



2/ 



h.)) 



'"Kh=J f ('"-^h.h)^^- 



h.h 



(4.9) 
(4.10) 



(4.11) 



To calculate the fljd-ng regimes common for a helicopter, we can assume a 
small value of the angle 3 . Then the flapping equation is simplified to 






(4.12) 



To determine the angle of attack of the blade section a^ , we examine the 

drawing (Fig. 2. 93) in a plane perpendicular to the 
blade axis (view along the arrow C in Pig. 2. 91). 

Figure 2.93 shows that air with a relative ve- /178 
locity U, directed at an angle $ to the plane of ro- 
tation, will flow over the blade section turned 
through an angle cp to the plane of rotation (cp being 
the blade pitch in the studied section) . The angle of 
attack of the blade section is equal to 




Fig. 2. 93 Speeds and 
Aerodynamic Forces of 
Blade Element. 



a,==cp-f tarT^ — =(p+tan"' =- ■ 



(4.13) 
(4.1!f) 

(4.15) 



The blade pitch in the examined section depends on the following: overall 



188 



pitch of the rotor 9o equal to the blade pitch in the section r = 0.7 at p = 
and without cyclic change of pitch; angle of twist Acp of the section relative 
to the section r = 0.7; flapping angle of the blade in the presence of a flap- 
piiag compensator; cyclic change of blade pitch. The sum of these terms is equal 
to 

tp = eo-t-Atp — ftp — 9iSin'^ — fljcos';* — y'(y„cosw!^ 4-v„sin«'J>), (4«l6) 

where 

9i and Q^ = conponents of cyclic change of blade pitch, with deflection 

of the automatic pitch control; 
v„ and v„ = conponents of elastic twist of the blade. 

The aerodynamic forces per unit length in the section r are determined by 
the coefficients Cy and c^ for the profile of the section under study,- taken in 
relation to a,. . Since, in detennining a^ , the induced velocity in the section 
was taken into account, the coefficients Cy and Cx are taken for a profile with 
infinite elongation. 

The Mach and Reynolds numbers in the section are 

(4.17) 



M = 


U 
a 


a 


-u= 


--M,U; 


Ub 

V 


= 


o^Rboj 

V 


bU = 


abr,-, = 

= "bM 

V 



Re=— = — "-dbU=^bM. (4.18) 

Since the Re for helicopters is rather high, the coefficients Cy and c^p of 
the sections will be considered (for sinplification) to depend only on the Mach 
number in the section. Therefore, the aerodynamic characteristics of the pro- 
file for each M are taken at Re corresponding to a given Mach and mean chord and 
flight altitude: 

Re^^f— ") *o.7M. (4.19) 

The lift and drag per unit length of the section will then be 

^_Xs^^±QU^bc. (4.20) 

dr 2 ^ 

QU^bc^., (4-21) 



dXsec.^ 1 



dr 



■xpi 



while their conponents directed along the axes relative to the rotor, i.e., /179 
the thrust dT and the resistance to rotation dQ, will read 

dr dr ^ dr (4.22) 

i^=^^^cos a* -^^^- Sin <». 
dr dr dr (4.23) 

189 



Substituting, into eqs.(4«22) and (4»23), the ejqjressions for cos $ and 
sin I from eqs.(4.24) and ik-25) 



cosO=^; (4.24) 

sin$=^ (4.25) 



and the expressions for — /""' - ^^'^ — '7^'°'— from eqs.(4'20) and (4-21), we 
finally obtain ^^ '^^ 



ar z. 

or, in relative quantities. 



dt 



'L={cM^^c^Uy)Ub; (4.28) 



dr 



dr 



'L={c,/J,-c,U,)Ub: (4.29) 



The antitorque moment of the blade per unit length, or the section torque, 
is determined from the formula (in relative quantities) 

^=^^[(F-7;jcosp+7,J. (4.30) 

After integrating the loads per unit length over the blade radius, we ob- 
tain expressions for determining the forces and torque of the blade. Since these 
quantities depend on the blade position in the plane of rotation (its azimuthal 
position ijf), they are given the subscript t • 

^,= f ^d7- (4.31) 



'h.h 

i ^^''; (4-32) 

'h.h 

^.=j' fK^-VJ^^^^+^ft.hl'^^" (4.33) 



'h.h 
'h.h 



'h.h 



The blade thrust is directed at an angle P to the axis of the rotor. Its 
projections onto the rotor axis and onto the plane of rotation are equal to 
t ^ cos p and t ^ sin g . 

190 



VcOSot. 




On mapping the TDlade forces in /180 
the plane of rotation onto the longi- 
tudinal and transverse axes of the 
rotor, we find the longitudinal and 
transverse forces of the blade 
(Big.2.94): 



h^ = 



^4, sin p cos iji -f ^^, sin ip; 
54,= — /,), sin p sin (j) — ^4, cos (j). 



(4.34) 
(4.35) 



Fig. 2. 94 For Determining Coeffi- 
cients of Longitudinal and Trans- 
verse Blade Forces. 



The confjonent t* cos p creates the 
longitudinal and lateral moments of 
aerodynamic forces m,, and m^ . : 

"^zj^^ = — ^4, cos p7j, J, cos <jj; ( 4 .36 ) 
'"'A+ = - ^+ cos p\^ sin if. (4 .37) 

In order to determine the blade forces and moments in a dimensionless form, 

i_ pa(cuR)^F and 

2 



the dimensionless coefficients must be multiplied by -^ 

-^ -i. pa(u)R)^FR. 

The instantaneous value of rotor forces and moments can be found by summing 
the forces and moments of all blades at a selected instant of time (one blade 



being at an angle i|i> the second at an angle ijr + 
+ 2 ■„ , and so on). 



2tt 



, the third at an angle i|r + 



The average per-revolution forces and moments created by the blade are 
equal to the integral with respect to ^ from eqs.(4«3l) - (4«37) divided by 2tt. 
On multiplying the result by the number of blades, we find the average forces 
and moments of the rotor per revolution. 

In a dimensionless form, the average per-revolution forces and moments of 
aerodynamic forces are determined by the e^^jressions 






2ic 



(4.38) 
(4.39) 
(4.40) 



191 





'"'A =i;r f '"'a /'I' = ~ 57^h.h f ^t cos p cos ^. rf^.; 
b 6 

2t ' 2« 



/IBl 

(4.41) 



(4.42) 



(4.43) 



A force equal to the sum of the inertia forces of blade flapping is trans- 
niitted through the flapping hinge to the rotor hub (Fig. 2. 95). Its projection, 
directed parallel to the rotor shaft axis. 



\= - J '« ^('-^hM)^''^ ^dr= -5^^0,2 ^cos p, 



hJi 



(4.44) 



creates longitudinal and lateral moments of the rotor 

27t 2i: 



2ic 2ti 



(4.45) 



(4.46) 



which should be summed with the moments of aerodynamic forces [see eqs.(4.42) 
and (4.43)]. 

We note that the integral expressions (4.42), (4.43)* (4.45), and (4.46) 

contain sin ^ or cos i|r , due to 
which the moments are created by 
the first harmonics of thrust and 
inertia forces. Therefore, the 
moments of the inertia force are 
greater in magnitude than the 
moments of the aerodynamic force, 
since the first harmonic of t* is 




Rotor axis -' f^ 



ju^r _. , > Since Tine iirsii narmonxc 01 t,^ 1£ 
df/^ ''^' small because of blade flapping. 



Fig. 2. 95 Generation of Rotor Moment 
Created by Inertia Forces of Flapping 
Motion. 



The lift and drag coeffi- /182 
cients of the rotor are determined 
by changing from the body-fixed 
system of axes to a velocity 
system: 



192 



/y = / COS a— A sin a; (4«47) 

^^=/sina + Acosa. (4-48) 



2. Method of Calculation 



The initial data for calciilation are the dimensionless rotor characteris- 
tics: geometric blade characteristics (variation in twist A9, relative chord t>, 
and profile over the blade length), load factor of the rotor ct, aerodynamic 
profile characteristics, stagger of the flapping hinges t^,^, mass and weight 

characteristics of the blade (-^ = P^°'''^ .. and ^f"'" ' ), coefficient of the 

flapping conpensator k. 

The operating regime of the rotor is given the following dimensionless 
data: angle of attack of rotor a, coefficients of velocity and lift V, ty, Mach 
number M,-,, deflection of controls h, T] (or angles 9i and 83 ) . 

The sequence of the calculation is as follows: In first approximation, the 
magnitude and distribution of the induced velocity v and the rotor pitch 9o are 
assigned. The induced velocity can be taken from eq.(3«46) or from e^qjerimental 
data. The rotor pitch is determined either by the Glauert-Iock theory with con- 
version by means of eq.(2.242) or is assigned arbitrarily (for exatiple, 9o = ty ) . 

Let us select the azimuth with which to begin the calc\ilation fo and the 

initial values of Po and Pq \fo^ brevity, we denote: p' = -g|-, P" = ^~| )• 

Usually, we take ^o = or to = 270°; Po and Pq can be determined by the 
Glauert-Lock theory or we can assume Po = Po = (which, for all practical 
purposes, does not lengthen the calculation since the natural oscillations of 
the blade decay rapidly) . 

At the initial azimuth we calculate |j,, \, U^, Uy , U, cp, a^ , and M at all 
radii, and then determine Cy and c^p from the graphs of the aerodynamic charac- 
teristics of the profiles. Next, ~z. and m^.j, are determined, and from the 

flapping equation (4.10) we find Pg . From Pg* Po» and Po we find, ty numerical 
integration, P and p ' at the next azimuth and continue the calculation in this 
sequence at other azimuths. 

In the method of calculation conpiled and programmed by M.N.Tishchenko, 
integration of the flapping equation of the blade with respect to azimuth is 
performed by the Euler method with conversion. From the values of Pj, P/, P'l 
at azimuth i|ri we find the preliminary values of P/+x and Pj + i at azimuth 
iItj+i from the expressions 

193 






where A^r = ijii+i - ^^ 



Then, from eq.(4»10) we calculate the preliminary value of the thrust /183 

moment coefficient relative to the flapping hinge (mj,,ii )i+i = f(Pi+i , 
Pi+lpr )• 

Furthermore, assijming first that in the section between azimuths 1^1 and 
'^i+i there is a uniformly accelerated motion with an average acceleration 

-i- (Pi + Pi+i ) and secondly that Pi+ipj. can be found from eq.(4«10) with re- 
spect to (m^.i, )i+i and Pi+i , the system of equations 



tpr 



P/+i = P/ + A4.; 

P/+i = P/-| Ai^; 



will yield, by the iterative method, the final values of ^(+1 and ?>i+i . Then, 
knowing p/+i and Pi+i , we calculate the final values of m^.h^^j and Pj+i . 

The calculations showed that, with this method, integration can be per- 
formed with an interval Ai|; = 12° . 

Integration of the loads per unit length over the radius, as well as forces 
and moments of the blade with respect to azimuth, is acconplished by the trape- 
zoidal method. For exanple. 



n 



1=2 

1 

1=1 



di \ '•*-^t_, 

* 2 



Here, k is the niimber of blade sections (r^ = 1, ri is the root section), 
and n is the n-umber of calculated azimuths- 

Using the described method, we then calculate one or two revolutions of the 
rotor and conpare the values of p ' and P" with those which had been at this 
azimuth in the preceding revolution. The obtained value of ty is conpared with 
that assigned. If these values do not agree within the stipulated acctiracy, 
then the difference ty - ty is used for refining the value of 9o and 

194 



calciilating another revolution of the rotor. 

The calculation is considered conpleted as soon as ty is equal to the as- 
signed value, to the required accuracy, and as soon as P and p' in the last and 
preceding revolutions coincide. 

As a result of the calculation, we determine the average forces and moments 
per revolution, the distribution of the section angles of attack, the thrust co- 
efficient, and the blade flapping angle, which are represented as Fourier series 
with an accin-acy to five terms: 

5 



^<^ = (+^{^„cosn^-{-t„sinn<^); (4.49) 

71=1 

5 

p=ao— 2(«/iCos«<l> + 6„sin«^). (4.50) 



Only the average induced velocity over the disk, determinable by eq.(3«46'), 
was taken into account in the calculations whose results are presented below /184 
in Subsections 4 - 7« The blade was considered to be absolutely rigid in bending 
and torsion. 

The integration interval was 12°, the number of calculated radii 12, and 
the accuracy within which ty, P', and P had to coincide was JAtylnax = 0.002; 
lAP'Uax = 0.002; |Ap|„ax = 0.002. 

At the blade tip, ■ _ and — ^ were calculated for Cv = and c,„ corre- 

dr dr ' ^ 

spending to Cy = 0. At sections r ^ 0.975* the calculation was made without ar^r 
corrections for taking tip losses into accotmt. 

The calculation time of one flying regime on a conputer performing 20,000 
operations per second, is 40 - 75 sec. 

3. Aerodyn^iiic Characteristics of P rofiles for Rotor Blades 

Below we give the aerodynamic characteristics of NACA 230 and NACA 00 pro- 
files and also of a high-speed profile suitable for use at the tip of helicopter 
blades. The first two profiles are taken at a relative thickness of 12^ and the 
last profile, of 9%. 

The aerodynamic characteristics of the profiles were obtained from test 
data on a rectangular airfoil model in a wind tunnel, with conversion to infinite 
aspect ratio and to full-scale Reynolds numbers taken for each Mach by means of 
eq.(4.19): 

Re=f— ") 6o7M = 20-10«M. 

The aerodynamic characteristics of the profiles in the angle of attack range 
from -2° to 15 and Mach numbers from 0.3 to 0.9 are given in Tables 2.6 - 2.8. 

195 



'^XP 



TABLE 2.6 
PROFILE NACA 23012 





Al\ 


—2 


1.0 
0.205 


3.5 
0.46 


7 


9 


11 


12.5 
1.365 


14.5 
1.525 


15 




0.3 


—0.085 


0,81 


1.035 


1.21 


1,525 




0.4 


-0.10 


0.20 


0.445 


0,80 


1.01 


1.20 


1.33 


1.42 


1,42 




0.5 


—0.085 


0.225 


0.485 


0.85 


1.0 


1.185 


1,24 


1.25 


1,245 


Cy 


0.6 


—0.085 


0.225 


0,485 


0.843 


0.94 


1.0 


1,03 


1.048 


1.05 




0.7 


—0.085 


0.245 


0.505 


0.715 


0.785 


0,837 


0.87 


0.91 


0,915 




0.8 


—0.065 


0.285 


0.43 


0.556 


0.625 


0.675 


0,715 


0.76 


0.77 




0.85 


-0.065 


a. 185 


0.30 


0.435 


0.490 


— 


— 


— 


— 




0.9 


—0.075 


0.09 


0.22 
0.010 


0.015 


0.018 


0,022 


0.029 


0,045 


— 




0.3 


0,008 


0.008 


0.05 




0.4 


0.008 


0.008 


0.010 


0.015 


0.023 


0,0355 


0.043 


0.07 


0.074 




0.5 


0.008 


0.008 


0.010 


0.019 


0.031 


0.0575 


0.0835 


0.121 


0.130 


••.rp 


0.6 


0.008 


0.009 


0.0135 


0,0365 


0.0765 


0.128 


0.167 


0.218 


0.230 


0.7 


0.009 


0.013 


0.0275 


0,09 


0.138 


0.181 


0.213 


0,254 


0.262 




0.8 


0.0125 


0.03 


0,067 


0.130 


0.177 


0.121 


0.253 


0.294 


0.304 




0.85 


0.028 


0.049 


0.080 


0.145 


0.185 


— 


— 


— 






0.9 


0.069 


0.08 


0,1075 


— 


— 


— 


— 


— 





TABLE 2.7 
PROFILE NACA 0012 



M 



—2 



3,5 



0,3 

0.4 

0.5 

0,6 

0.7 

0.8 

0.85 

0.9 



0.3 

0.4 

0.5 

0.6 

0.7 

0,8 

0.85 

0.9 



—0,185 

—0.18 

—0.215 

—0.215 

—0.235 

—0.245 

—0.19 

—0.08 



0.0095 

0.0095 

0.0095 

0.010 

0,010 

0,0245 

0.0415 

0.069 



0.085 

0.095 

0,10 

0,11 

0.11 

0,135 

0,095 

0,02 



0.007 

0.007 

0.007 

0.007 

0.0085 

0.016 

0.036 

0.069 



0.32 

0.335 

0.355 

0.375 

0.395 

0.40 

0.29 

0.14 



7 


9 


11 


0,645 


0.835 


1.02 


0.665 


0,85 


1,035 


0.71 


0,915 


1.08 


0.75 


0,91 


0,94 


0.735 


0,81 


0,84 


0.57 


0,65 


0,72 


0.50 


0,61 


0.71 


0.40 


0.56 


0,70 



12.5 



14,5 



15 



0.009 


0.0125 


0.0165 


0.021 


0,009 


0.0125 


0.0165 


0',021 


0.009 


0.013 


0.0185 


0.031 


0.0105 


0.021 


0,039 


0,074 


0.0185 


0.061 


0,0955 


0,135 


0.046 


0.095 


0.131 


0,1675 


0.061 


0.1065 


0.141 


0,180 


0.0795 


0.118 


0,149 


0.187 



1.155 


1.34 


1,39 


1.175 


1.25 


1.25 


1.1 


1.1 


1.1 


0.95 


0.96 


0.965 


0.860 


0.863 


0.865 


0.765 


0.765 


0,75 


0.0240 


0.029 


0.034 


0.0245 


0.061 


0.080 


0.051 


0.106 


0.126 


0.1095 


0,171 


0.186 


0. 1675 


0,211 


0.221 


0.195 


0,2285 


0.236 


— 


— 


— 



mi 



196 



TABI-E 2.8 
HIGH-SPEED PB3FILE 





M\ 
0.3 


—2 
-0.065 


1.0 

0.235 


3.5 
0,485 


7 

0.835 


9 

1.035 


11 
1.18 


12.5 


14.5 


15 




1.165 


1.115 


1.1 




0.4 


—0.065 


0.23 


0.485 


0.835 


1.035 


1.10 


1.09 


1.06 


1.05 




0.5 


—0,065 


0.245 


0.50 


0.86 


1.015 


1.015 


1.0 


0.99 


0.99 


'> 


0.6 


—0.065 


0.26 


0.53 


0.90 


0.98 


0.96 


0.965 


0.96 


0.96 




0,7 


—0.07 


0.30 


0.60 


0.96 


0.96 


0.935 


0.935 


0.95 


0.95 




0.8 


—0,07 


0.36 


0.63 


0.81 


0.87 


0.87 


0.89 


0.935 


0.945 




0.85 


—0,12 


0.325 


0.55 


0.77 


0.86 


0.86 


— 


— 


— 




0.9 
0.3 


-0.165 
0.008 


0.175 
0.007 


0.46 
0.009 


0.815 
0.011 


0.012 


0.0245 


— 


— 


— 




0,065 


0.12 


0.133 




0.4 


0.008 


0.007 


0.009 


0.011 


0.012 


0.055 


0.0975 


0.142 


0.15 




0.5 


0.008 


0.007 


0.0095 


0.0125 


0.046 


0.093 


0.13 


0.1765 


0.1885 


Cjp 


0.6 


0.008 


0.007 


0.010 


0.025 


0.060 


0.110 


0.1475 


0.195 


0.205 




0.7 


0.008 


0.0075 


0.015 


0.061 


0.10 


0.143 


0.175 


0.195 


0.221 




0.8 


0.0125 


0.012 


0.037 


0.092 


0.128 


0.165 


0.194 


0.2125 


0.2415 




0.85 


0.021 


0.026 


0.053 


0.11 


0.15 


0.19 


— 


— 


— 




0.9 


0.044 


0.04 


0.069 


0.131 


— 


— 


— 


— 


— 



In the calculations whose results are given below, when M < 0.3 we took /1B6 
the profile characteristics for M = 0.3, whereas when M > 0.9 the coefficients 
Cy and Cjtp were determined tj linear extrapolation with respect to M = 0.85 and 
M = 0.9. If the angle of attack of the blade sections varied within 72° to 180° 
and -7° to -180°, the characteristics of all profiles were determined regardless 
of M from Table 2.9. At angles of attack from 15° to 72° and from -2° to -7°, 
a linear interpolation was made between the corresponding values of Cy and c^p . 



TABLE 2.9 



a" 


72 

0,35 
1.1 


105 

—0.33 
1,1 


170 

—0.62 
0.04 


—170 

0.77 
0.15 


—105 

0.27 
1.08 


—85 

—0.2 
1.08 


—70 


—7 


Cy 


-0.32 
0.87 


—0.62 
0.04 



FigiH-es 2.96 and 2.97 contain graphs of the coefficients Cy and c^p as a 
function of or at all three Mach values. 

For selecting a profile at a small portion of the blade (for exanple, at 

.197 



the tip portion), graphs of the aerodynamic characteristics of the profiles as 
a function of angle of attack are more characteristic than the profile polars, 
since the angle of attack of the examined blade section depends little on Cy of 
this section and is determined mainly by the flight regime (ty, V, a) and the 
blade shape. Consequently, when the profile is changed, the angle of attack of 
the section is not changed (bearing in mind that ao of the profiles differ by 
less than 1 to 1.5°). To select the profile for a blade as a whole, the profile 
polars or the graphs of the aerodynamic characteristics of profiles constructed 
as a function of o? ~ cxo are m.ore characteristic . 

In Figs. 2.9s and 2.99 we have constructed the graphs of a^r and cv^^ as a 
function of the Mach number (a^r is the critical angle of attack at which Cy of 
the section begins to decrease or a pre-vious increase stops; Cy = Cy ^ ; a""" is 
the angle of attack at which a marked increase in c^p begins, oviing to flow 
separation or owing to wave drag) . Since the Mach number of the blade section 
is approximately equal to 

M«MoC7:t = Mo(P+Fsin\J))=rMo + Myj sini|j, 

the graphs in Figs.2.9S and 2.99 give the value of a at which an increase in 
profile drag begins and separation phenomena appear as a fimction of a combina- 
tion of Mf 1 , Mq, r, and i|; for the blade section. These graphs will be used in 
Subsection 8. 

The graphs in Figs. 2. 96 and 2.97 indicate that, for M = O.3, the NACA 23012 
profile has Cy^^^^ = 1.53, whereas the thinner high-speed profile has Cy^j^^^ = 

= 1.18 at O'er = 11°. For the latter, a steep increase in c^p begins as soon as 
a > 10.5°. The NACA 0012 profile has Cy„^^ = 1.4. 

At M = 0.6 - 0.7 and at average angles of attack, the profile characteris- 
tics are close together, whereas at M = 0.9 the high-speed profile is more ad- 
vantageous, having the lowest value of c^p at small angles of attack and a 
normal slope of dependence of Cy on a . 

The results of calculating rotor profile losses for different profiles are 
described in Section 3*4: At low Mq, the rotor with symmetric NACA 0012 profile 
and, in certain cases, rotors with the thin high-speed profile on the blade tip 
have profile losses several percentages lower than rotors with other profiles; 
at high Mq, the rotor with the high-speed profile at the blade tip definitely 
has the upper hand. 

The maximum permissible value of the lift coefficient, in terms of flow /lA 
separation (see Subsect.7) of a rotor with a NACA 23012 profile is by a factor 
of 0.01 - 0.02 larger than for a rotor with a high-speed profile at the blade 
tip. 

The aerodynamic characteristics of profiles should include corrections to 
account for the quality of manufacture and design features of the blades. The 
profile drag as well as the quantity Cy^^^^ are influenced by the flexibility 
and roughness of the surface (fabric skin or plywood cover, spacing of ribs), 
by the presence of projecting parts especially near the nose of the profile 

198 



ZMZ 



Cy 




^ 




M= 


0.3 




/■ 


^ 


Cy 
















/ 


/ 


' 


W 
0.5 




M=0.9 


/ 




— 




^/ 


f} 


y^ 


'^ 






/ 


10 










-4 


r 


-^ 




/ 


V 








■ / 




d 










0.5 































5 oc- 


A 






















5 




10 




OC" 


Cy 


// 






M=0.6 








W 


2 


.:^ 


5^ 


^ 


— 


=^ 




/ 










0.5 


/ 


























// 


y 

















NRCR2W12 

NflCR OOIZ- 

^ *mm M^High- speed 
pro fi I € 



10 



oC 



Pig. 2. 96 lift Coefficient of Different Profiles. 



Cxp 








o- 




■ 












/ 




~ NHLH 6o\Jfc 




0.20 


NnCAOOIZ 

High-speed 


JA^ 


^ 










/ 


^ 






/, 


/ 




profil e 




/ 




/ 


/ 


/ 




'^ 
_ 


t^' 


^^' 




/ 
^ 






/ 


f 


^ 




/ 


• / 


/ 


f 




y 






1 


'/ 






/ 


0.10 




A 


/ 


/ 


J 


/ 


<: 


4 


/' 

y 


/ 


f 


/ 




/ 




^^ — - 


/ 




> 






*ii,^ 


7^ 




' 




^ 


" 


s= 


^ 


:r^. 






■^, 






--^ — 








■""^ 






■ ' ' 


f 




i — 



M-OJS 



'M'0.3 



10 



15 oC 



Fig. 2. 97 Profile Drag Coefficient of Different Profiles. 



199 



(de-icing system, rivets), leakage in the joints of the blade segments, and 
local deviations from the theoretical section profile. 

On the basis of calculations, it is recommended to increase the Cjjp values 
of the profile, obtained from model wind-tunnel tests, by Ac^p equal to: 
for blades with a nose in the form of a continuous spar of metal, 
plastic, or wood and with rigid shanks: 0.0 - 0.001; 
for blades of segments with metal skin and ribs: 0.0015 - 0.0025; 
for blades with veneer or fabric covering: 0.0025 - 0.005. 

One or another value of Ac^p is selected from the indicated interval, de- 
pending on the quality of blade manuf actiore . 



75 



70 







































- 




























"«: 




^ 




































































' 






















- 
















■^ 


























-. 












^ 






























-- 










■ 


NACmjOIZ 



























NACfiOOlZ 

—' — — High-speed " 
profile 


















































- 


1 1 1 1 I 1 1 1 1 1 











0.3 



OM 



0.5 



0£» 



«" 




\ 










- 




- 




•- 


- 


- 




\ 




nI 








\ 




.^ 








^ 




\ 




s, 








^S, 


\ 












\ 


\ 


\ 






















5 


- 


















r 


V 








5 






















NKKZZOn 

NKHOOU 

High-speed 

profi le 


s 












\, 


^^ 


N 












^ 












v 


s 


X 

N 




p 




' 




\ 


N 




0.3 








0. 


5 


- 




av 

























Fig. 2. 98 Critical Angle of Attack of Fig. 2. 99 Angle of Attack of^- at which 
Profiles as a Fiinction of M. the Profile Drag Begins to Increase, 

as a Function of M. 

4. Distribution of Aerodynamic Forces over the Rotor Dis k 

Only the average induced velocity over the disk was taken into account in 
the calculations whose results are described in Subsections 4-7; the error 
thus introduced into the total _average characteristics of the rotor revolution 
at large and average values of V is small. The blade was considered to be abso- 
lutely rigid in bending and twisting. Calciilations show that flexural deforma- 
tions have practically no effect on the average aerodynamic per-revolution 
characteristics of the rotor whereas partial deformations, if the blade is in- 
sufficiently rigid, do have a noticeable effect. Preassigned torsional deforma- 
tions can be taken into account by substitution into eq.(4.l6).. 

The calculations were performed for eight variants of geometric blade /190 
characteristics, given in Table 2.10, with the following initial data: a = 



Y 



= 0.091; k = and 0.4; -3^ = O.9 and 1.2; I 



b. h 



=61 =02 =0. 



In this Subsection, we will examine the distribution of aerodynamic forces 



200 



, # 



TABLE 2.10 



No. of 

Blade 

Variant 



III 

IV 



VI 



VII 



VIII 



Bl adj3 Sh ap e 




Trapezoidal twisted with high-speecf 
p ro f i J. e 



Rectangular twisted with high-speed 
pro-file 



Rectangular twisted 



Rectangular twisted with 
symmetric profile 



Trapezoidal flat with high-speea 
p ro f i 1 e 



Hectangular with high-speed 
profile and increased twist 

Expanding with high-speea 
p ro f i 1 e 

Rectangular with larger portion 
of high-speed profile 



1.0 
1.0 



1.0 



Geometric 
Twist 
from 

r=0 to 

r=l,"deg 



Bl ade Profile 



From Root 

to ^=0.75 



1.82 0.625 2.9 



1,0 1.0 



1.0 
1.0 



1.82 0.625 



1.0 1.0 



0,59 1.176 



1.0 



1.0 



1.0 7 

1.0 7 

1.0 7 

2,9 

1.0 15 

0.5 7 



23012 



From root 

to 7=0.65, 
23012 



From /■=0.75 

to F-0.85 



23012 Transitional 



From_r=0,65 

to r=0.75, 

Transitional 



From r=0.85 
■to 7=1 



23012 Transitional High-speed' 

23012 Transitional High-speed 

NACA 230.12 
NACA 0012 

23012 Transitional High-speed 



High-speed 



Transitional I High-speed 



From /■=0.75 

to r=1 

High-speed 



O 
H 






over the rotor disk. The asstuiption of constancy of induced velocity and the 
absence of blade deformations leads to errors in calculating the forces distri- 
buted over the rotor disk; consequently, the data in this Subsection are only 
approximate* 

Let us examine rotors vri.th blades of variants I and II in two helicopter 
flying regimes: one close to horizontal flight and one close to autorotation of 
the rotor; both regimes are taken at equal lift coefficients ty = 0.16, dimen- 
sionless velocity V = 0.3, and Mq = 0.6. The results of rotor calculations in 
these regimes are given in Table 2.11. 



TABLE 2.11 



ty = 0.16; V = 0.3; Mo = 0.6 



Characteri sties 



Horizontal Flight 



Rectangul ar 
Blade 



a" 
X 

eo 

tx 
h 

K 

H 
as 

h 

h 

^1 






k=0 



—9.4 

—0.0610 
7.820 

—0.0095 
0.00849 
0.0168 
0.0997 
0.0973 
0.0398 
0.0069 

—0.0025 
0.0004 
0.00015 

-0.0059 
0.0219 
0.0309 

-0.0143 
0.0062 
0.0033 



A=0.4 

—9.4 

—0.06103 
9.957 

—0.0101 
0.008698 
0.0162 
0.09667 
0.09535 
0.003355 
0.006043 

—0.003146 

—0.000548 

—0.0002637 
0.003116 
0.0249 
0.03466 

—0.00963 
0.0062 
0,0089 



Trapezoidal 
Blade 



k=0 

-9.4 

-0.0610 
8.0320 

-0.00795 
0.00796 
0.01815 
0.0949 
0.108 
0.0408 
0.0078 

-0.0024 
0.0005 
0.00027 

-0.0062 
0.0233 
0.0377 

-0,0149 
0.0078 
0.0055 



*=0 

-10.3 

-0.065 
8.45 

-0.01 
6.0086 
0.0186 
0.0958 
0.1096 
0.0405 
0.0076 

-0.00276 
0.000628 
0.000177 

-0.00623 
0.0227 
0.0364 

-0.0167 
0.00859 
0.00407 



Autoro tation 



Rectangular TrapezoidaJ 



Blade 



Blade 



k=0 
1.4 

-0.0048 
3.576 
0.0168 
0.000475 
0.0129 
0.0926 
0.06938 
0.0367 
0.00559 

-0.00203 
0.0003 
0.00014 

-0.0053' 
0.0289 
0.0249 

-0.0120 
0.00427 
0.00348 



k=OA 

1.4 

-0.0048 
5.62 
0.0172 
0.000365 
0.01327 
0.09247 
0.07166 
0.00857 
0.00515 

-0.00269 
0.000457 
0.000175 

-0.0054 

0.0285 
0.0267 

-0.0153 
0,00623 
0.00317 



A=0 



1.4 

-0.0048 
3.550 
0.0180 

-0.00015 
0.0140 
0.0877 
0.0772 
0.0368 
0.0062 

-0.0012 
0.0003 
0.0003 

-0.0058 
0.0312 
0.0300 

-0.0138 
0.0051 
0.0049 



202 



Table 2.11 shows that, in horizontal fUght, the flapping motion of the /191 
blade is greater than in autorotation and that it is greater for the trapezoidal 
than for the rectangular blade. A conparison of the characteristics of rotors 
with rectangular and trapezoidal blades shows that, at equal a, ty, and V, the 
rotor with trapezoidal blades has larger 9o , h, ai, bi and a smaller absolute 
propulsive coefficient t^; at equal tx, ty, V (see the column with o' = -10.3°) 
the rotor with trapezoidal blades has a more negative angle of attack, and the 
difference in the quantities Sq , h, ai , bi increases even more . This is re- 
sponsible for the change in balancing characteristics of a helicopter when the 
trapezoidal blades are replaced by rectangular types (for exanple, there is a 
decrease in deflection of the automatic pitch .control forward, owing to a de- 
crease in the longitudinal force H and in the coefficient c i) . 

















«f 






















^ 


\ 

^=0 


— 




N 


X 






W 


ex 


\ 
















-tf'\ 










(/'=18lf 


- 


- 












\ 














oc—9A°^ 




.\ 












-- 








<> 


^5>^ 








- 
















^^^ 


■^ 


s.. 





















r'0375'0.75 0.55 \25 

(/f'90 



0.25 



0.55 0.75 0.975 r 




0.55 0.75 0S75 r 



Fig. 2. 100 Angles of Attack of Sections as a 
Function of Blade Radius . 



Figure 2.100 shows graphs of the variation in angles of attack of a rotor 
with rectangular blades with respect to blade radius at four azimuths: 0, 90, 
180, and 270° . The solid lines refer to horizontal flight and the dashed lines 
to autorotation. 

In horizontal flight, the angles of attack are negative at the blade root, 
at t = and 270° . At azimuth t = when the flapping angle is small, the 
vertical velocity conponent equal to about V(q' + 3 ) - v = V(Qr + a.^ - a.^) - v 
(Fig. 2.101) is large and directed downward, due to which the angle of attack at 
the blade root is small or negative at this azimuth. At azimuth i|; = 270° the 
roots are close to the zone of the backwash and are washed backward and upward. 
Therefore, they have large negative angles of attack. 



In the middle and tip sections of the blade, the angles of attack are 



203 



iiiiiiim III 




*^=/ffO 



Fig .2. 101 Conponent of Mr Veloci- 
ties Normal to Blade Axis at Azi- 
muths if = and ■}/ = 180°. 



greatly influenced by the peripheral ve- 
locity of flapping. This increases the 
angles of attack at azimuths of 270° 
and 0°, where the blade is shifted down- 
ward and decreases them at azimuths of 
90° and 180°. The geometric twist of /192 
the blade reduces the increase in angles 
of attack toward the blade tip at azi- 
muths of 270° and 0° and decreases them 
even more at azimuths of 90° and 180° . 
The angles of attack are negative at the 
blade tip at azimuth of 90°. 



In autorotation, the angle of attack 
of the rotor is positive so that the angles of attack of the blade root sections 
have a large positive value. At the blade tips, the angles of attack of the sec- 
tions are less than in horizontal flight, since in autorotation the rotor has a 
small pitch and less flapping motion. 

The distribution of the section angles of attack over the entire rotor disk 

with twisted rectangular blades is 



Rectangular blade 



It IJ, 



= -9.4 




Fig. 2. 102 Distribution of Angles of 
Attack over Rotor Disk (Horizontal 
Flight). 



illustrated by the graph in 
Fig. 2. 102 (horizontal flight). The 
hatched circle in this diagram is 
the zone of backwash, along whose 
boundaries the section angle of 
attack is close to ±90° . Regions 
with negative section angles of 
attack are also shown by hatching. 

Figure 2.102 shows that, in 
the zone bounded by azimuths of 
270 - 300° and relative radii 
0.7 - 1.0, the angles of attack 
reach a maximal value (for un- 
twisted blades, the angles of at- 
tack are maximijm at r = l.O). 
This region, in which flow separa- 
tion_takes place on increase in ty 
and V, has a noticeable effect on 
rotor operation as a whole. In /193 
autorotation the separation region 
is located in the root portion of 
the blade at azimuths of 200 - 300°. 



Calculations show that the 
maximum angles of attack of a trape- 
zoidal blade are substantially- 
larger than those of a rectangular 

blade. For a rotor with a flapping conpensator, the maximum angles of attack 

at ijt = 270° decrease somewhat. 



204 



The linear thrust is extremely unevenly distributed over the radius and azi- 
muth (see Fig.2.2S), which is responsible for the occiirrence of the large vari- 
able bending moments of the blade. It follows from Table 2.11 that, for the 
rotor with trapezoidal blades and for the rotor with a flapping conpensator, the 
variable portion of thrust increases. The rotor with a flapping conpensator has 
a larger fourth haraionic. These peculiarities of rotors must be taken into ac- 
count when estimating vertical vibrations of helicopters. The magnitude of the 
variable thrust conponent depends on the rotor characteristic |j, (or V): the 
larger \i,, the larger the variable conponent. At cos ijf - t^ , the coefficient of 
the first harmonic of thrust is very small; consequently, at small spacing of 
the flapping hinges the aerodynamic moment of the rotor relative to the trans- 
verse axis m. 



^j [eq.(4«43)] need not be taken into consideration. 



0.02 



0.01 



-.11 '' ^ 


-r -^''li^ 




— K'O ^\ J n i\ 1 f 


"\Tv l5^?^^^- 



inn 



^^iy30o (p° 




Fig. 2. 103 Torque Coefficient of Blade 
vs. Azimuth (Hoiizontal Flight). 



Fig. 2. 104 Acceleration Moment in 
Blade Section dQ = dXp ^ 



X cos $ - dY„ 



see. 

sin $ < 0. 



The torque coefficient of the blade mt 



varies greatly with the azimuth 



(Fig. 2. 103). At azimuths where the section angles of attack increase (quad- 
rants III and IV), m^ markedly decreases. In an autorotation regime, m,., is 

negative in the quadrants III and IV. This is due to the fact that, at these 
azimuths, the blade sections have large positive inflow angles $, as a result of 
which the projection of the lift of the blade section is directed forward and 
produces an accelerating moment (Fig. 2. 104). Thus, it is obvious that, in 
forward flight in autorotation regime, decelerating moments are produced in the 
quadrants I and II and accelerating moments in quadrants III and I? (during 
vertical descent of a helicopter in an autorotation regime, the decelerating 
moments are produced by the tip sections of the blade and the accelerating 
moments by the root sections). 



It should be noted that a very large variable torque conponent, in a rotor 
with the usual stagger of drag hinges (T^,^ < 0.05), produces a small (within 
1°) flapping motion relative to the drag hinges, since the eigenfrequency of 
blade oscillation is by a factor of about 4 lower than the rotor ipm, i.e., the 
frequency of change of m^x. 

In Section 2, we noted that, at equal ty, t^, and V, the quantity mt does /194 
not depend on the anplitude of cyclic change of blade pitch, whereas a change 



205 



of m^ with respect to azimuth does. Actually, as shown in Fig. 2. 103, for a 

rotor with a flapping conpensator m^. differs with respect to magnitude and 
phase. ^ 

5. Aerodynamic Characteristics of Rotor 

The aerodynamic characteristics of a rotor are presented in graphs in the 
form of the dependence t^ = f(mt) with parameters ty, a at Mf^ = const (or V = 
= const), Mq = const. These graphs are convenient for determining the coeffi- 
cients tx, mt, and the angle of attack q- from values of Mfi, V, Mq, ty known 
from an aerodynamic calculation. 



-(7.(7/ 



-0.03 




Fig. 2. 105 Aeroaynamic Characteristics of Rotor 
(Mfi = 0.0975; V = 0.15; Mq = 0.65; a = 0.09I). 



Such graphs for a rotor with rectangular twisted blades with a high-speed 
profile at the tip (variant II of blades) are shown for Mq = O.65 in 
Figs. 2. 105 - 2.109. 

We see from the graphs that the dependence of the propulsive coefficient t, 
on the torque coefficient m^ is practically rectilinear, with the exception of 
near-separation values of ty at negative t^ , where the rate of increment of mt 
increases owing to an increase in profile losses. In these cases, curves with 



206 



different ty become nonequidistant. The interval between the ciirves increases 
with increasing ty, which can also be attributed to an increase in profile 
losses with increasing ty . 

Curves corresponding to very small values of the lift coefficient /I95 
(ty < 0.1 - 0.08) closely approach or intersect the curves corresponding to 
large values of ty . This means that a decrease in rotor thrust coefficient (for 
e3caiiple, when using a wing on a helicopter) to ty = 0.08 and less it not recomr- 
mended since, in this case, the propulsive force of the rotor does not increase. 
The \:pward deflection of the curves with small ty \;pon a decrease in m,. shows 
that, at small ty, the rotor is not in an autorotation regime. 



"•X 
O.OJ 



0.02- 



0.01 



-001 



-0.02 



-0.03 




Fig. 2. 106 Aerodynamic Characteristics of Rotor 
(Mfi = 0.13; V = 0.2; Mo = 0.65; cr = 0.09I). 



The advantages of the described graphs conprise: sinple shape of the 
ciirves, facility of interpolation -upon variations in the coefficient ty, and 
the possibility of using them for different solidity ratios (see Sect .6). With 
the use of these graphs for calculating balancing and stability, we can deter- 
mine the coefficients t and h by means of the conversion formulas (3»15) and 
(3.17). 



To determine rotor pitch in the calculation, we use the dependence 9o = 



= f(mt) with the_parameter ty 
parameter Sq at V = const, Mq 



at V = const, M© = const, or ty = f(Q?) with the 
= const. The graphs of the relation ty = f(a) are 



207 




/196 



-OM 



Fig. 2. 107 Aerodynamic Characteristics of Rotor 
(Mfi = 0.195; V = 0.3; Mo = 0.65; ct = O.09I). 



ix 


H 


> 


"^ 


\, 






— 












S 


•X 




\ 








- 
























X: 


:^ 


N 


V 




li m 






■^ 


^ 


> 




\ 


•v 








ty = 0.06^ 
iy-0.08-^ 
ty=0.J0-^ 
ty-OM^ 


y 


> 




V 






•ck.* 


^ 


1 


^5 


% 








- 


S 



























0.0025 








0.0050 




1 


^ 


-•0.0073 "1 


k 


>i 


0.0 


w 


0.0125 


mt\ 
























n/i- 


^ 


WN 


X 


































ty=O.OB^ 

tx=o.ds^ 


/ 




^ 


^ 


f^ 


V 


\ 




-^^ 
















































'5^ 


^ 


N 


\ 


^ 




^ 


X 


'< 


*-^ 










0.01 










o a = 5° 
A 0° 
• -5° 
A -70° 
X -75° 
P -20° 




















^ 


^ 


sT 


\ 




ki-'.. 




f 








































^:^< 


^ 


'«? 




^ 


































iy-a04' 
ty-B.06^ 
ty=OW^ 






>i 


S 


s 




\ 








^-. 


























^ 


y 






^ 


^ 


\ 


V 


X 










































V 


^ 




5» 


\ 


^ 


pi\? 






















































oa 


— 


?5 


■ 










































^ 


^ 



Fig. 2. 108 Aerodynamic Characteristics of Rotor 
(M,i = 0.26; V = 0.4; Mo = 0.65; a = O.O9I). 



208 



shown in Figs. 2. 115 and 2.116. 

The aerodynaniic characteristics of rotors with blades of different shapes 
are not presented here since, with the accepted assunption that regardless of 
blade shape the induced velocity is distributed unLfomly over the rotor disk, 
the difference in the coefficient mt at given ty, t^, V, Mq is determined /197 
entirely t^r the difference in nipp . Therefore, our conclusions concerning the 
effect of blade shape obtained in examining graphs of mp,. in Section 3.3 remain 
unchanged. 



coos 



'0.005 



-0.010 



-O.O'S 




Fig. 2. 109 Aerodynamic Characteristics of Rotor 
(Mfi = 0.325; V = 0.5; Mq = 0.65; o = O.091). 



6. Aerodynamic^ Qha^ ^acteristics of Rotor in Autorotation Regime 

The graphs of the characteristics of a rotor of the variant II in an auto- 
rotation regime - polars ty = fCtj,^ ), rotor performance K^, pitch Bq^ , and angle 
of attack a^ - are shown in Figs. 2. 110 - 2.113. 

The graphs indicate that the rotor performance is lower than that of a wing 
(for more details on rotor performance see Sect.? of Chapt.II). At small V, 
autorotation of the rotor takes place at large positive angles of attack. Upon 
an increase in rotor rpm (of Mq) the rotor performance drops, the pitch de- 
creases, and the angle of attack increases. 

It is known that autorotation of a rotor is possible in the absence of 
forward speed and at any low flying speed. Therefore, the minimal permissible 
speed of an autogiro or helicopter on engine failure is not determined ty flow 
separation at the wing, control lability loss or spinning, as would be the case 
in regular aircraft, but by the permissible vertical rate of descent. In verti- 
cal descent, the rotor develops approximately the same drag as a plate (cx « 
« 1.28) with an area equal to the rotor disk area, with the vertical speed of 

209 




7198 



Fig. 2. 110 Polar of Rotor in Autorotation Regime (a = 0.091). 




0.16 



0.18 



0.03 0.10 0.12 O.lt 

Fig. 2. Ill Rotor Performance in Autorotation Regime (o = 0.091). 



210 



i:m. 




Fig. 2.112 Rotor Pitch in Autorotation Regime. 




Fig. 2. 113 Angle of Attack of Rotor in Autorotation 
Regime (a = O.O9I). 



211 



I 



descent of an autogiro being 



V^ 



1 / =3.5 l/^ . 

y 112-1.28 qq/^F y A 



In autogiros, the value of p was small and their rate of vertical descent 
was low. 

Figure 2.104 shows that the acceleration (negative) moment in an auto- /200 
rotation regime is created by the projection of lift; consequently, autorota- 

tion is not possible at small rotor 



[tr^lmin 



0,10 
































- 


-^- 
















































^ 




- 




. 
















- 






005 



















m 



0.20 



0.30 



0.W 



Fig. 2. 11!)- Minimim lift Coefficient in 
Autorotation Regime. 



Hft. Figure 2. lift- gives a graph of 
the minim.um Hft coefficient in 
autorotation (ty )„!„, which is 

either the autorotation limit or the 
value of ty at which autorotation is 
generated at very large angles of 
attack and negative pitch. 

In a helicopter with a large 
wing, the rotor lift markedly de- 
creases during autorotation, and, 
since the rotor cannot have a very 

small coefficient ty, autorotation occurs at a lower rpm than in helicopter 

regimes . 

To estim.ate the influence of the geometric blade characteristics on the 
autorotation regime of a helicopter, we present the following data. 

At optim-um gliding speed (V = 0.2) when Mq = 0.7 for a helicopter with a 
rotor not having a high-speed profile at the blade tip, the vertical rate of 
descent increases by 1.7 m/sec and the flight-path angle 9fi.p by 2°; the pitch 
should be 0.5° smaller. The angle of attack increases by 1.8° while the pitch- 
ing moment is retained (A«? - Aa + A9fi.p - O). 

At Mq = 0.6, the deterioration in autorotation characteristics is less by 
a factor of 2 - 3. 

Change-over to trapezoidal blades reduces the vertical rate of descent by 
0.65 m/sec and the flight-path angle by 0.8°. 

7 . limit of Permissible Helicgpjter Flight_Re^ime s 
"CFIow Separation limitj 

As shown in Subsection l+, a rotor with flapping hinges has areas with large 
angles of attack of the blade sections. In helicopter flight regimes (hori- 
zontal flight, gain in altitude) these are located at the blade tip at azimuths 
of 270 - 300° and in the autorotation regime, at the blade root at azimuths of 
200 - 300°. 



212 



An increase in lift coefficient ty caiises formation of a zone of super- 
critical angles of attack on the rotor. 

Fiirthermore, the rotor disk contains zones of high and supercritical angles 
of attack, at sites where the blade passes close to vortices shed by the pre- 
ceding blades. Here the blade enters a region of high ttpward-directed local 
induced velocities causing an increase in the angles of attack of individual 
sections . 

As soon as the zones of sipercritical angles of attack become large, the /201 
rotor characteristics change noticeably: The dependence of ty on the pitch and 
upward angle of attack becomes nonlinear, and the coefficients of flapping, 
longitudinal and lateral forces, and profile drag of the rotor all increase. 

The limit of permissible regim.es with respect to flow separation conditions 
is determined by the magnitude of the rotor lift coefficient ty characterizing _ 
the average level of the section angles of attack, by the velocity coefficient V 
characterizing the degree of nonuniformity of distribution of the section angles 
of attack over the rotor disk, by the rotor angle of attack determining the 
character of the distribution of the section angles of attack, and also by the 
blade shape and the separation characteristics of its profile. 

From the expressions for the coefficient ty and V 

y G 



y 



1 1 

it is obvious that, on a decrease in rotor rpm and an increase in flying speed 
and altitude, the coefficients ty and V will increase so that the helicopter may 
enter a flow- separation regime. The phenomenon associated with flow separation 
at the rotor blades can be stopped rapidly by decreasing the pitch, increasing 
the rotor ipm, and reducing the flying speed. 

Deep penetration into the flow-separation zone sometimes ends in catastrophe 
for the helicopter. One of the most inportant problems of selecting the heli- 
copter parameters in designing and determining its flight characteristics is to 
ensure absence of flow separation in all permitted flying regimes. Owing to the 
possibility of entering the flow-separation zone, the maximum flying speeds and 
altitude are limited on helicopters and any decrease in rotor rpm below an 
established liaiit is inpennissible. In order to avoid flow separation at high 
flying speeds, a wing is installed on helicopters to reduce rotor lift. 

Flight tests show that flow separation manifests itself by an increase in 
blade stresses and in blade hinge moments, increase in helicopter vibrations, 
imbalance of the helicopter, and deterioration of control 1 ability . Consequently, 
the manifestations of flow separation differ widely and are conplex for deter- 
mining the limit of separation by calculation. Flight tests and wind-tunnel 
tests of rotors yield insufficient data for establishing the overall limit of 

213 



flow separation. Therefore, the limiting values of the lift coefficients ty 
obtained by calculation are given below. 

Calculated graphs of permissible values of thrust coefficients are given 
in the literatiire (Ref .20, 24). In the first of these papers, the limiting 
fUght regime is that regime at which the average lift coefficient of the blade 
at azimuth ij; = 270° becomes equal to the maxim-um lift coefficient of the pro- 
file Cy . In the second paper, the criterion of flow separation in helicopter 

flight regimes is taken as the equality of the angle of attack at the blade tip 
at azimuth ijt = 270° to some critical value a^^ : Q?or ~ l2° in a regime corre- 
sponding to the start of separation phenomena, and a^^ = 16° at the limiting 
flight regime with a large separation zone. 

A shortcoming of both methods is that one does not know how to select Cy 

or Qfgj. for a blade with a set of profiles. Furthermore,, the degree of nonuni- 
formity of distribution of the angles of attack over the rotor disk depends /202 
on V; at ty = ty^^, and at large V, the zone of increased angles of attack occu- 
pies a smaller portion of the disk than at small V. • Therefore, the appearance 
of sipercritical angles of attack at large 7 has a_less pronounced influence on 
the change in rotor characteristics than at small V; this is not taken into ac- 
count in the method presented in the second paper (Ref .24)- 



















^ 










































• 


















%-ir - 














- 
















































^^ 
















^ 


































^ 


^, 


lc(-/l- 


-_■ 








." 


^ 


r-" 


'^ 








W 


0.i 




^ 


^■ 














^ 






























- 


- 




^^ 




^ 








-- 




- 
















00*^1 


















/^ 


y 








- 


















-^ 


( 
















^ 


y 
















^ 


y' 
















y 


^ 




- 


' 


















X 


X 












.\ 


(f 


^ 


y 








0.15- 








X 














«o^ 


^ 






















^ 




( 












X 


y 
























^ 


^ 




























^ 




- 






^ 


^ 


















































/ 














































/ 





























































































































































































































-10 



-15 



-10 



a' 



Fig. 2. 115 Change of Coefficient ty as a Function of 
Angle of Attack and Pitch of Rotor 
(V = 0.2; Mo = 0.7). 



In the separation limits constructed below, it is assimied that the permis- 
sible magnitude of the coefficient ty is the value at which the character of the 
dependence of ty on the angle of attack and pitch of the rotor begins to change. 
Such limits are constructed for rotors with different geometric characteristics 

214 



/203 





Cy 






1 I , 




^ 




g-l5 -- 


__ — — ^ 














__^ — -"^ .0 ^---' 


ttZy __ 


u 




^ ■^ " 6iiJ^^ ^ 


^^^ 




" A^- 


-^r^ 'ycr -^'^ 


^ 


_ 


H-W|a5° 


-t=^ I.^- 




— 


r >^ 


" 1/./5 




/ 


/ / 






/] [ 


i^jfl / • 






/ 






/ 


/ 


/n/> 






/ 






r 


'/V 






z 


<?ktt/ 






!~ 7 


> 






L /I 


/ 






/ 


/ 


" a(/J 




V 


'' I _ - 






/ 






\ T / 


^ 






/ 







-ZQ 



'-15 



-w 



-5 



5 cL' 



Fig. 2-116 Change in Coefficient ty as a Function of 
Rotor Angle of Attack and Pitch (V = 0.4; Mq = O.?)- 





i 


^ 










°f 


h,\h^ 


^ 


A 


- 






■4 


y 










/ 










/ 


0.2' 


o.op^ 


/ 








/ 




i 


/ 




/ 








/ 


/ 




















0.7 


0.025 










5. 




c 
o 

a 








/ 


^ 








^^ 


'-- 


-^ 




X 




y 


^ 




" 








--' 




/ 


7 














1 — 

















-/ 




m 
nj 


L 


1. 


/ 

4— 

e 
o 

o 

1. 


/ 




--| 


777 


p'*. 


777 


-| 


/ 


/ 






/' 








1 




0.0 
/ 

o.c 


3 
2 


- 


— 










a 










^ 


- 


- 


- 


— 










1" 





-20 -15 -10 -5 5 oc* 



-20 -15 -10 -5 



5 oc° 



Fig. 2. 117 Dependence of Coefficients 
of Longitudinal and lateral Forces h 
and s and of Flapping a^ and hi on 
Rotor Angle of Attack (V = 0.4; 
Mo = 0.7; eo =11°). 



Fig. 2. lis Dependence of Coefficients 

of Torque m^ and Profile Power lEp^ on 

Rotor Angle of Attack (V = 0.4; 



M„ 



0.7; eo = 11°). 



215 



on the basis of calculations lay the method presented in Subsection 2. 

Figures 2.115 and 2.116 give graphs for the dependence of the coefficient ty 
on the angle of attack and pitch of a rotor with blades of the variant II, at a 
flapping conpensator coefficient of k = 0.4. Figiire 2.115 indicates that, at 
V = 0.2 when ty reaches a certain value, the increase in ty practically stops. 
The coefficient ty has a limiting value which it cannot exceed at any a and 9© . 
Thus, because of the small V, the flow separation extends over a large zone and 
there is a marked change in characteristics. Since the incipient deflection of 
the cvccve from the linear segment is not well-defined, we will use, at V = 0.2 
and for ty , a value of ty less than the maximum by an amount corresponding to 
Aff = 2° (Aty « 0.01). 

At V = 0, the maximum possible value of ty is taken for t^j. (Sect .8). At 
large V (see Fig. 2. 116), the increase in ty with respect to a markedly slows 
down at some value of ty . The value of ty at which the curve deviates from the 
linear law by Ao? = 0.5° is taken for ty . 

Figures 2.115 and 2.116 indicate that the quantity ty at given V and Mq 
depends little on the rotor angle of attack. 



Figures 2.117 and 2.118 give graphs for the dependence of the coeffi- 
cients of longitudinal and lateral 
forces h and s, flapping a-^ and bj,. 



/204 



0.35 


\ 


s 


- 










- 










- 


\ 


\ 


S 


s. 


^ 


^ 


=^ 








s 






\ 




V 


s 






\ 




\ 


-* 






0.30 




\ 


ii 




X 


V 














\ 






s- 


\ 


















\ 
















^ 


"^ 




s 










\ 
\ 














<c- 


> 


\ 






\ 


0.25 


— 


- 


- 




\ 




























0.20 




















































m.i 















torque m^, and profile power mpr on the 
rotor angle of attack. These coeffi- 
cients also change when ty = ty : The 

forces of the rotor and the flapping 
motion of the blades increase backward 
and to the side of the advancing blade 
(^If = 90°), and the profile power coeffi- 
cient increases markedly. The variable 
portion of rotor thrust, i.e., the 
second and higher harmonics, also in- 
creases substantially. 

Thus, at V > 0, the value of t„ 

' ' 'or 

is smaller than the maximum possible 
values of ty ; however, it can be assumed 



0.10 



0.20 



0,30 



that, as soon as 



ty = 



the ■ above- 



Fig. 2. 119 Dependence of ty on V 
and Mq (Rectangular Twisted Rotor 
with High-Speed Profile at the 
Blade Tip). 



mentioned phenomena associated with flow 
separation will become manifest. 

The_curve for the dependence of 
ty on V and Mq is plotted in Fig. 2. 119* 

It is obvious that the quantity ty de- 



At small and medium V, tj 



creases greatly ipon an increase in_V. 
decreases with increasing Mq, whereas at large V 



the effect of Mq is insignificant. 



216 



Z20i 



0.35 




Fig. 2 .120 Dependence of ty on V of Rotors with 
Blades of Different Shapes (Mq = 0.4). 





V. 

\ 


\ 




N 








- 














1 




































- 








































































\ 


\ 


n 


^ 








































\ 


\ 


1^ 




















U-Zi 


\ 


N 


i^ 
























\ 


s 


'N 
































\ 




^ 


V 






























\ 




^ 


^ 






























\ 


s 




^ 


*^s 














X 


^ 


"^^ 


^^ 


































^s^~ 








^ ^^ 


































"^■-v^ 


^I 


nfiT 
























■^/ 



0.10 



DM 



0.3 



O.fV 



Fig. 2. 121 Dependence of ty on V of Rotors -with 
Blades of Different Shapes (Mq = 0.?). 



217 



Figures 2.120 and 2.121 give graphs of t„ for blades with different geo- 

' c r 

metric characteristics. The rotor with rectangular blades of MACA 230 profile 
(variant III) has the largest value of ty . The same rotor with a highr-speed 
profile at the tip (variant II) has values'" of ty smaller by 0.01 - 0.02. An 

r 
increase in geometric twist of the blade increases ty by approximately 0.01 

(variant IV). The rotor with trapezoidal blades (variant l) has the smallest 
value of t„ 

'or 

The graphs of ty are approximate and obtained by calculation, but they 

do permit the helicopter designer to determine the limit of safe flying speeds 
before conducting special helicopter flight tests. Flight tests show that we 
can obtain a slightly larger value of ty than the calculated values of ty 
This is e^^ lai ned by the fact that our accepted ty are smaller than the maxi- 
mum possible values of ty, and also by the fact that factors that increase Cy ^ 

were not taken into account in the calculations, namely effect of centrifugal /206 
forces on the boundary layer and unsteady flow through the rotor blades. 

8. Distribution of Profile Losses Over Rotor Di_sk. 
Dependence of Profile Losses on Aerodynamic 
Characteristics of Blade Profiles 

In Section 3*3 we examined graphs of the coefficients of rotor profile 
losses. Let us define the extent of influence of aerodynamic characteristics 
of the blade profile, peripheral speed, and blade shape on the distribution of 
profile losses over the rotor disk and their total magnitude. 

The required power of the rotor, referred to all-up weight, is proportional 
to mtM§ [see eq.(5.l6) in Chapt.IIl]: 

^ — »<3 a , — ,,3 

— = /ny.Mo , = const, /re. Mo , /■ ^-, \ 

G * 75iCyMl ' f (4.51) 

where 

/rejMo = const2-|-/re^rMn. (4.52) 

Thus, the required power of a helicopter at given Mf ^ , H, p = -y-, c^ = 

^°g,^ - is determined by the quantity ffippMo calc\JLated at values of CyM| and C^M^ 

corresponding to the given quantities. For exanple, a helicopter has a load per 
square meter of the rotor disk area of p = 35 kg/m^ and a parasite drag coeffi- 
cient of "cx = 0.0075; the calculated (operating) flight regime is V = 275 km/hr 
at a height of H = 1000 m. Under these conditions, the dimensionless coeffi- 
cients of a helicopter are equal to 

CyN[l = -^-— = 0.00545; 
218 



/207 




-JSM 



Fig. 2. 122 Angle of Attack at Blade Tip Section as a 

Function of M, for Three Values of Peripheral 

Rotor Speed (Mq). 



CrnM 




300 </>' 



Fig. 2. 123 Profile Losses in Blade Tip Section as a 

Function of Azimuth Position of Blade, for Three 

Values of Rotor Peripheral Speed (Mq ) . 



219 



V 



Mfi =—=0.227; 
CxI^l=c^N[}i =0.00039. 



Let us give the resiilts of calculations pertaining to the tip section of 
the blades r = 0.975. Fig. 2. 122 is a graph for the angle of attack change of the 
tip section of a rectangular blade (variant II, a = 0.091) vri.th respect to azi- 
muth, plotted as a function of M of the section for three values of the peri- 
pheral speed (Mq). We see from the graphs that the section has large angles of 
attack at small M and small negative angles at large M. By means of these 
graphs, it is easy to determine the location of flow separation zones at inr- 
creased profile drag. For this, it is necessary to plot the cvocres of a^^ and 
0?^'" for the profile of the investigated section (see Figs. 2. 98 and 2.99) on the 
graphs in Fig. 2. 122. It is obvious that, for a highr-speed profile and for Mq = 
=0.7 (u)R = 235 m/sec, V = 0.325, ty = 0.1228) the maximum angles of attack are 
low (1.5° lower than the critical values) but at azimuths i|f = 35 - 1^0° deep 
penetration into the region of. high c^p takes place. At Mq = 0.655 (ujR = 
= 220 m/sec, V = 0.347, ty = 0.14) the maximum angles of attack are close to 
critical and there are two zones of high profile drag: at azimuths i|r = 55 - 120° 
and at azimuths t = 270 - 0° when of = 10 - 5° and M = 0.41 - 0.62. When Mq = 
= 0.61 (u)R = 205 m/sec, V = 0.373, t, 
= 250 - 350° ^ 
profile drag. There are no increases of profile losses at large M at azimuth /208 

i|r =90°. 

The permissibility of deeper penetration of the tip section into the flow- 
separation zone from the point of view of rotor behavior as a whole is charac- 
terized by the graph in Fig. 2. 119* In conformity with this graph, a flight 
regime with Mq = 0.61 is pennissible. 

As shown above, the required power of a helicopter and the profile losses 
of the rotor are determined ty the quantity mp^Mf which, for the examined sec- 
tion, is equal to 



^y = 0.l6l) the tip section at azimuths i|r = 
"'° penetrates into the flow-separation zone and into the zone of high 



Mo 

dr 2n 






(4.53) 



Figure 2.123 gives graphs for the product c^pM^ plotted against azimuth. 
The integral of eq.(4.53) is equal to (Table 2.12): 





TABLE 2.12 




Mo 


0.7 


0.655 
0.0058 


0.61 


dmpr 
dl ^» 


0.0073 


0,0056 



220 



Consequently, the greatest profile losses in the section under study occur 
at Mq = 0.7 and the smallest losses, at Mq = 0.61. At Mq = 0.655, the profile 
losses are somewhat greater than at Mq = O.6I, but local separation phenomena 
are absent. 

Mow let us assume that we were to change the profile in this section. Its 
angles of attack would then change slightly, while the zone of flow separation 
and high profile losses might change substantially. The curves of a^^ and a* 
are also plotted in Fig. 2. 122 for the NACA 23012 profile. Obviously, at all Mq 
the section would have no separation zones but would have a large zone of high 
profile losses at azimuths i|f = 280 - - 170° . Especially high will be the 

at 



losses at i|( = 90°, where M is greater than M^p by 0.1 - 0.2. 
all ■ii , the ipm of a rotor with this profile could be reduced. 



Since a^ < a^ 



Thus, the graph in Fig. 2. 122 gives the optimum dependence of 0-0^ and a'' 
M of the profile, for the section under study at one of the design flight 
regimes . 



on 



M, 



For exaiiple, the rotor profile losses would decrease if the profile had 



0.9 at Of « while retaining a""" = 7*5 - 5.5 




Fig. 2. 124 Angle of Attack of Blade Tip 
Section as a Function of Mach Number. 



at M = 0.5 - 0.6. Then the 
best rotor rpm woiild corre- 
spond to Mq =0.7. The thin 
symmetric high-speed profile 
has a high value of M^r at 
Of « 0, but a low value of o?"" 
at M = 0.5 - 0.6. 

The profile with M„ = 
= 0.8 at ex = and cv"''" = 
= 14 - 7° at M = 0.4 - 0.6 
woiild be suitable for the ex- 
amined flight regime for the 
case of Mo = 0.-61. A highly 
concave profile with a small 
relative thickness does have 
such characteristics; however, 
its use would considerably in- 
crease blade torsion and con- 
trols stress of the helicopter. 



In selecting the profile, it must be considered that the dependence o?, = 
= f(M) will be different in different flight regimes. Figure 2.124 gives a 
graph for a regime corresponding to flight close to the dynamic ceiling: Mfi = 
= 0.122; CyM| = 0.0103, CxM§ = -0.00012, Mq = 0.7. We see from Fig. 2. 124 that, 
in this regime, the p2?ofile losses are very high. In conformity with Fig. 2. 119, 
this regime lies at the boundary of flow separation. In hovering flight near /209 
the ground, at a lower peripheral speed, the examined section will have a^ = 
= 2.7°, M = 0.65. 

By suitable selection of blade shape, a certain influence can be exerted 
on the change in angles of attack at the blade tip with respect to azimuth and a 
better combination can be obtained of the dependence a^ = f (M) with the profile 



221 



characteristics. As typical exanple. Fig. 2.125 gives a graph of a^ = f(M) for 
the tip section of blades of the variants I (trapezoidal), II (rectangular), 
VI (rectangular with increased twist) and 711 (eD^anding) at ty = 0.12, t,,. = 
= -0.013, V = 0.4, Mo = 0.7. 




Fig. 2. 125 Angle of Attack of Blade Tip Section as a 
Function of Mach Mimiber, for Rotors with Blades 
of Different Shapes. 



Figure 2.125 shows that the trapezoidal blade, for which a reduction of 
chord at the tip (b^ip < l) leads to a decrease in profile losses, has the 
largest angles of attack at M = 0.4 - 0.?. The expanding blade, at these values 
of M, has angles of attack by 1° lower than those of the rectangular blade. The 
blade with increased twist has angles of attack by 1.7° lower at all azimuths 
than those of other blade variants. 

The integral in eq.(4.53) should be calculated to obtain a quantitative /210 
estimate of the effect of a change of blade shape and profile. 

It is clear from the foregoing that the azimuthal distribution of profile 
losses in each blade section depends on flight regime, peripheral speed, and 
blade profile. Main enphasis should be placed on selecting a suitable profile 
in the blade tip sections, where the largest profile losses occur. For illus- 
tration. Fig. 2. 126 shows the distribution of profile losses over the radius of 
a blade of variant II for Mf^ = 0.227, Mq = 0.655 at four azimuths, as well as 
the distribution of the average circumferential profile losses over the blade 
radius. Figure 2.126 indicates that about 35^ of profile losses are accoiinted 
for by the tip portion of the blade from r = 1.0 to r = 0.9. 

Section 5. Vortex Theory of Rotor 

1. Problems in Vortex Theoiy 

The main problem in the vortex theory of a rotor lies in the determination 

222 



of aerodynamic loads on the blade, -with consideration of the nonuniform induced 
velocity field. 

The solution of this problem perniits: 

1. Eefining the aerodynamic characteristics of the rotor. These refine- 
ments are less inportant for the single-rotor helicopter and more inportant for 

multirotor helicopters, where the 
mutual induced effect is very strong 
and has a substantial influence on 
their fUght characteristics. 

2. Determining both the constant 
and variable aerodynamic loads of /2ll 
the blade and from these loads calcu- 
lating the oscillations of the blade 
and its deformations. Without con- 
sideration of the nonuniform induced 
velocity field, ar^r determination of 
the variable aerodynamic loads on the 
blade in a mmiber of flight regimes 
is quite inaccurate. Therefore, the 
rotor vortex theoiy must introduce the 
conponent of blade oscillations and 
the determination of variable stresses 
into the calculation, i.e., into the 
stress analysis of the blade. 

OnHy by means of the vortex theory 
is it possible to explain such phe- 
nomena as the marked increase in vari- 
able loads on the blade and vibrations 
of the helicopter in low-speed regimes 
as well as the appearance of local 
flow separation zones at medium and 
high speeds. 
Fig. 2. 126 Distribution of Profile 

Losses over Blade Radius. In low-speed regimes, the induced 

velocity field is particularly non- 
uniform. This leads to the occurrence 
of appreciable variable aerodynamic forces acting on the blade. The blades 
begin to vibrate at increasing anplitude. Extensive variable stresses are set 
vp in the blades. The variable forces transferred from the blades to the hub 
lead to increased vibrations of the entire helicopter. The explanation of this 
phenomenon is possible only by making use of the vortex theory. 

At high and mediimi flying speeds, a phenomenon is observed which we can 
call induced flow separation. This phenomenon is a consequence of large induced 
velocities arising in the region of vortices shed from the blade tips. When the 
next blades pass below these vortices, appreciable surges in aerodynamic loads 
and, in certain regimes, even flow separation are created. This phenomenon was 
partially described elsewhere (Ref .17; and has been confirmed in flight tests. 

223 




A no less iaportant problem of the vortex theory is the determination of 
the induced velocity field caused ty the rotor in the stream flovd.ng past the 
helicopter and its individual conponents in flight. 

The character of flow past the wings of a helicopter, its fuselage, and 
stabilizer is largely determined by the velocity field induced by the rotor. 
The occurrence of induced velocities leads to additional downwash and to a change 
in the true angles of attack of the lifting elements and hence in the forces 
acting on all outer surfaces of the helicopter conponents. Therefore, to study 
the flow around these parts, it is necessary to determine the induced velocities 
at various points in the space surrounding the helicopter. 

Thus, the vortex theory permits determining the induced downwash in the 
region of the helicopter wing and its stabilizer and hence the aerodynamic 
forces acting on them. Therefore, the theory also introduces the following com- 
ponents into the calculations of aerodynamic characteristics: balancing of the 
helicopter, characteristics of its stability, and controllability featiires in 
which these forces play a substantial role. 

There are other phenomena for whose calculation the vortex theory is used. 
A sufficiently detailed description of all these phenomena is possible only in 
special works. Therefore, in this Section we will give only a brief accoiint of 
the most important elements of the vortex theory, without detailed substantia- 
tions . 



2. Theoretical Schemes for the Vortex Theory of a 
Rotor with a Rinite Number of Blades 



In the vortex theory, the rotor is replaced by a system of botind and free 
vortices. This system can be represented by a vortex sheet covered with horse- 
shoe vortices (see Fig. 2. 128). The /2l2 
segments of these eddies located at 
the blade are known as bound vortices. 
Depending on the purpose of the calcu- 
lation, we can use schemes in which 
the blade is replaced either by a 
bound lifting vortex (lifting-line 
scheme), or by a bound vorticity layer 
(scheme of a lifting- vortex surface). 
In the latter case (Fig. 2. 127), the 
blade is replaced by a system of boiind 
vortices distributed over the blade 
chord with some strength Yt, so that 




Fig. 2. 127 Flow around Blade Profile 
in Scheme of lifting-Vortex Surface. 



6/2 



r= Jy,^ dx. 



-6/2 



(5.1) 



where 



224 



r = velocity circulation over a contour enconpassing the blade section 
(Fig.2.128); 
Y^o = circxolation per unit length of the bound vertices distributed over 
the profile chorxi. 



The scheme of a lifting-vortex surface more accurately reflects the 
phQTsical pattern of flow around the blade but is more conplex in calculations. 
Therefore, to sinplify calculations, the lifting-vortex surface is often re- 
placed iDy a lifting- vortex line. In determining the induced velocities at a 
sufficient distance from the blade, this sinplifi cation does not produce exces- 
sive errors in the results and 
therefore is often used in calcu- 
lations. The induced velocities 
close to the blade must be deter- 
mined by the scheme of a lifting- 
vortex surface. 



During operation of the rotor, 
the conditions of flow around the 
blade at different radii are dis- 
similar. Thus, the magnitude of 
circulation of the bound vortices 
varies over the blade radius . A 
change in circulation is accom- /2l3 
panied by the formation of so- 
called longitudinal vortices (see 
Fig. 2. 128). The longitudinal 
vortices are a continuation of the 
boiind vortices located on the 
blade and form the tails of horse- 
shoe vortices extending to in- 
finity. 




Bound vortices 
V 



Longi tudina 
vor t ices 



Transverse 
vort ices 



Fig. 2. 128 Diagram of Formation of a 
Vortex Sheet in Circulation of Flow 
about the Blade. 



The strength of the longitudinal vortices should be equal to the change in 
circulation of the bound vortices over the blade radius: 



- — 



(5.2) 



where 



Yio = strength of longitudinal vortices per unit length; 
r = total circulation of the bouud vortices. 



If the circulation of the bound vortex changes in time, also transverse 
vortices will trail from the blade. The circulation of transverse vortices is 
equal to the change in circulation of the bound vortices with respect to time 



'tr 



or 
dt ' 



(5.3) 



where Ytr is "the circulation of transverse vortices shed "by the blade in unit 
time. 



The strength of the transverse vortices per unit length can be determined as 

1 dr 



\ U dt ' 



(5.4) 
225 



where U is the velocity con^Donent of relative flow, normal to the blade axis . 

Transverse vortices, just as bound vortices, form part of the horseshoe 
vortices and merge along the edges with the longitudinal vortices. As a conse- 
quence, the circulation of the longitudinal vortices is variable over their 
length and changes by the magnitude of circulation of the transverse vortices 
merging there. 

Under conditions of axial flow past the rotor, the circulation in the blade 
section T remains constant in time. Therefore, a vortex sheet consisting only 
of longitudinal vortices will be shed by the blade. Their strength proves to be 
constant over the length of the free vortex. 

3. Form of Free Vortices 

Under flying conditions, the free vortices shed by the blade are carried 
away from the rotor at a rate equal to the relative velocity of the flow passing 
through the rotor. These velocities, generally speaking, are different at dif- 
ferent points of this flow. Therefore, the free vortices trailing from the 
blades are carried away from the rotor at different rates . As a result, a 
rather complex vortex system will exist downstream of the rotor, which, more- 
over, is continuously being deformed due to the mutual interference of the vor- 
tices. At some distance from the rotor, the vortex sheet begins to be dislodged 
and finally loses its original form. 

It is extremely difficult to take into account deformations of the system 
of trailing vortices. Therefore, in theoretical methods of calculation few at- 
teDpts have been made to take these deformations into consideration. Usually, 
most authors assime that the free vortex sheet is carried away from the rotor /ZlU 
at a constant rate equal to the mean velocity of flow through the rotor. 

The con5)onents of this velocity with respect to the coordinate axes, re- 
ferred to the peripheral speed of the blade tip cuR, are usually taken as equal 
to |j, and Xq av (^oav being the average velocity of the flow of the stream along 
the axis of the rotor, referred to cuR) . 

The average flow velocity A-oav is determined by the well-known fonnula: 

^oav= (^fena — ;^ 



With such an assunption, the trailing vortices are arranged over a downwash 
spiral surface. The longitudinal free vortices are located along downwash spiral 
lines, whereas the transverse vortices are arranged over the radial generatrix 
of this spiral siirface. Therefore, as applied to a helicopter rotor it is 
preferable to divide the trailing vortices into spiral and radial rather than 
into longitudinal and transverse, as is done in the airfoil theory. 

All free vortices trailing from the blades are located within an inclined 
cylindrical surface resting upended on the circumference of the rotor. The 

226 



"vortex system enclosed ■within this siirface is usually called a vortex column or 
a vortex cylinder. 

Let us derive the equation of the line along which are located the free 
vortices shed from the blade at an arbitraiy radius p . This Une coincides with 
the wake of the blade in the flow passing through the rotor. Neglecting devia- 
tion of the blade from the plane of rotation, the coordinates of this Une (see, 
for exanple, the coordinates of point A in Fig. 2. 130) can be written as follows: 

x= — ecosO — (a;?(i})o — "y);! 

2=— QSind; \ . - 

where 

-" < ^ < \|fo ; 
< p < R; 
^Q = azimuth of the blade at the instant of time in question; 
■& = azimuth of the blade at the instant of time of shedding the vortex. 

All confutations given below will be based only on such a form of the free 
vortices. We will disregard refinements introduced by consideration of their 
deformation. 

4. Determination of the Induced Velocities by the 
giot-Savart Formula 

If the form of the free vortices is known, the Biot-Savart formula can be 

used for determining the induced velocities. This formula permits determining 

— ♦ 
the elementary velocity dv induced at the point M by a vortex element of length 
dS (Fig. 2. 129). In vectorial form, the formula can be written as 

^-£%?, (5.6) 

where ^ 

dv = vector of the elementary induced velocity caused at the point M by 
a vortex element of length dS; 
r = circulation of the vortex; 

dS = vector of the vortex element; 

t = vector proceeding from the point M where the induced velocity is /2l5 
_^ calculated to the locus of the vortex element dS; 
I = |t| = distance from the point M to the vortex element dS. 

The direction of the vector dv is perpendicular to the plane formed by the 
— * — > 

vectors dS and t . 

To determine the induced velocity from the total vortex, eq.(5.6) must be 
integrated with respect to its length: 

227 



1) = 



J 4it 



13 



(5.7) 



Having taken such integrals over the length of all vortices, located both 
on the blade and shed by the rotor, we can obtain the total induced velocity at 
any point of space around the helicopter. 



5. Use of the Biot-Savart Formula in Devej -oping the 
Vortex Theory of a Rotor 

Equation (5«7) can be used as basis of the rotor vortex theory. For this, 
it is necessary to deterniine the induced downwash in the rotor plane and take 

this into consideration when determining 
the true angles of- attack of the blade 
sections. After this, the loads on the 
blades can be determined l^r formulas of 
the type 




r=.- 



CyQbU\ 



(5.8) 



Fig. 2. 129 Diagram for Calculating 
Induced Velocities by the Biot- 
Savart Formula. 



where the value of Cy is taken with re- 
spect to profile wash, for angles of at- 
tack calculated with consideration of in- 
duced downwash. 

To determine the induced downwash 
on the rotor blade it suffices to calcu- 
late only the axial (parallel to the 
rotor axis) conponent of the induced ve- 
locity Vy . Then the downwash angle can 
be approximately determined by the formula 



Aa„ = 



U^ 



(5.9) 



where Ao-y is the change in angle of attack of the blade element due to- downwash. 

Accoi*ding to the rules of vector analysis, the projection of the product /2l6 

—4 — » 

of the vectors dS x I onto the y-axis can be calculated in the following manner: 



(dS X Oy = dS^l^ - dSJ^, 



(5.10) 



where dS^ and dS^ are projections of the vector dS, while tx ^"^ ^z ^•J'e projec- 
tions of the vector t onto the x- and z-axes; the direction of the x- and z-axes 
is shown in Fig. 2. 130. 



228 




d-+Zn 



' oav 



Fig. 2. 130 Diagram for Calculating Axial Components 
of Induced Velocities. 



Correspondingly, the axial conponent of induced velocity can be obtained 
by a formula analogous to eq.(5-7): 






(5.11) 



Usually, the induced velocities are represented as the sum of three comr- 
ponents: 



'" = 'ybo -T-^sp +'yrad. 



(5.12) 



where 



v,,o = induced velocity from bound vortices; 

Vgp = induced velocity from spiral vortices; 

Vpad = induced velocity from radial vortices. 

Let us construct the general formulas for calculating the axial conponents 
of induced velocities, without l±miting ourselves to the case y = (y being the 
coordinate of the point at which the induced velocity is calculated) . 



229 



6. Axial Component of Induced Velocity from Bound Vortices /2l7 

Boimd vortices have a circiiation equal to F,^ and are located along the 
rotor blades. The subscript N denotes here the numeral of the rotor blade. We 
will consider that N = 0, 1- 2, ..., z,, - 1, where z,, is the number of blades 
in the rotor (see Pig.2.13o5. 

The values of dS^ and dS^ entering eq.(5.1l) for the bound vortices are 
equal to 



^^^^'^bo ==~dQsmi/ff, J (5.13) 



where ijr^ is the azimuth of the blade with the numeral M. 

Substituting eqs.(5.13) into eq.(5.1l) and taking the integral of eq.(5.1l) 
over the length of all rotor blades, we obtain 



^=-6-' R 



N=0 

where 



■"bo = 2 ]r^f<j,dQ, ^ 



4it [is ^'^ ' /3 ^^J 
The values of l^ , t^ and 1 entering into K^ are determined by the formulas 



ljc= r cos <^ ~ Q cos <^f/, ' 
4 = rsint|j — Qsin^;v. 



(5.15) 



where r is the radius of the circumference passing through the point at which 
the induced velocity is calculated, while 



'^N=% + — N. (5.16) 



Here, iItq is the azimuth of the blade with the numeral W = 0. 

Substituting the values of t^ and l^ into the formula for K^, we obtain 



1 rsin(^ — <); ) 

Aa' = 

4--t /•-) 



7. Axial Component of Induced_Velocity from Spiral 
( Longitudinal )jyortices 

In determining induced velocities from spiral vortices, we must sum the 
230 



velocities induced by -vortices trailing from different radii of the blade. As 
in the determination of velocities from bound vortices, we must define here the 
total velocities induced by vortices shed from all rotor blades. 

To use eq.(5.1l) for this purpose, it is convenient to divide the vortex 
sheet into strips joining the vortices shed from a portion of the blade of 
length dp . Then the circulation of the vortices enclosed in this strip, in ac- 

cordance with eq.(5.2), is equal to — dp. 

To determine the values of dS^ and dS^, we differentiate eqs.(5«5)- Then 
(see Fig. 2. 130), 

Using eq.(5.1l), we obtain /2l8 



-" -S| I^^-^^''*' 



Ar=0 »=— « 

where 



f.,» = . . . . . 

(5.18) 



^^=^T[7i-^^°^^+-&(e^^"'+t^^)]- 



The values of t^ , Ij , and I entering into L^ are determined by the formulas 
(see Pig. 2. 130): 



/^=— ^ A//V + '^ cos tj) — Q cos S, 

An 



'^Oav 



/^ = rsintjj — QslnO, 



(5.19) 



where 

The value of Hn, representing the distance along a normal to the plane of 
rotation, from the rotor to the point of the vortex sheet (see Fig. 2.130), can 
be expressed in terms of the azimuth of the vortex sheet j? by means of the 
formula 

///v = ('l'o-»+^^^)U,/?. (5.20) 

where N = 0, 1, 2, ..., z^ - 1. 



231 



8. Axial Component of Induced Velocity from 
Radial (Transverse) Vortices 

The circiilation of radial vortices shed from the blade in unit time is 

equal to - /* • dt. In the vortex wake, these vortices are located over the 
ot 



radial generatrix. Therefore, 



{dS,\^^= -dQ sin ^. J (5.21) 



Using the same approach as above, we obtain 

2 -I R 



^m<i= - 2 [ [ ^ ^^"^^ "'^' 



where 



N=0 /=— oo 



(5.22) 



= — [--^sln&+-^cos»l. 
4it L '3 ' /3 J 



The values of l^ , l^ , and t entering into M^ are determined by eqs.(5-l9). 

Integration with respect to t in eq.(5.22) can be replaced by integration 
with respect to ^, bearing in mind that ?? = ilt^j + cut. Then, 

z, -1 R t/v 



"'«'-- 2 H ^^'"'»'"- (5.23) 



A'=0 »=— 00 



The functions K^, L^,, and M^ entering into eqs.(5'14), (5.18), and (5*23) /219 
will henceforth be called the induction coefficients of the vortex element dS 
and a point with the coordinates r, ijr, and y. 

9. Integrodiff erential Equation of the Vortex Rotor Theory 

To determine the aerodynamic load on the blade profile, the induced down- 
wash must be determined from all trailing vortices and bound vortices of all 
blades, with the exception of the blade in question, since this vortex partici- 
pates in the formation of lift expressed by Joukowski's formula: 

T = QUr. (5.24) 

In other words, the downwash from the bound vortices on a blade with the 
numeral M = must be determined by calculating the induced velocity v^o by a 
formula differing from eq.(5»14) in that the term with N = is absent: , 



232 



^=' « (5-25) 

We then equate the lifts determined by eqs.(5.8) and (5 •24)- This jrLelds 



^^-^'^ybU. (5.26) 



If we limit oursel-ves to flight regimes in which it can be assumed that 
Cy = c^or and U = Uj^ and if we represent the angle of attack a as 

a = a>o + Aa„, (5.27) 

(where §o is the inflow angle to the blade profile which woiild be present in the 
absence of induced downwash Aay), and AoTy is expressed in conformity with 
eq.(5-9)» then eq.(5.26) can be written in the foim 

^-~^lbU^%+\clbv^. (5.28) 

Substituting here the value of Vy detemiined by eqs.(5«l2), (5.14), (5 -IS), 

and (5«23) and taking into account the refinement (5«25), we arrive at the in- 

tegrodifferential equation analogous to the basic integrodifferential equation 
of finite wing theory (Ref.28): 






z. -1 R 

Si 

,N=\ 



rpjKNdQ^ 






N=0 »=— ~ /V=0 »=— «o 



(5.29) 



It is necessary to note that the functions Uj^, $o , T^ , ^ ^ , -^^ > K^, 

op OJ? 

Lfj, and M^ entering here represent functions of the radius and azimuth of the 
blade. The function $o also depends on the flapping motion of the blade which, 
in turn, is a function of aerodynamic loads and hence of the values of T^ . 
Therefore, the integrodifferential equation (5.29) must be solved together /220 
with the equation of the flapping motion of the blade. 

It is not possible to suggest any general method of solving this equation. 
In each individual case, the method of solution most siaitable to the particular 
case is used in relation to the method of deterTiiining the induced velocities. 
As an exanple, we mention the method presented in Section 8 of Chapter I in 
Vol.11. 

For a solution, the method of successive approximations is occasionally 

233 



used, which involves assimi i ng at first that the induced velocity Vy is constant 
over the rotor disk and Tq is calculated in the first approxmation. In that 
case, the terms in brackets in eq.(5.29) are calculated and the new value of Fq 
is found; the procedure is continued in this manner until the solution converges. 

However, it must be borne in mind that convergence of this method is en- 
sured only in individual particular cases and therefore must be separately 
checked each time. 

Many authors, considering one or another method of successive approximations 
and believing it possible to restrict the process to the first approximation, 
only give a method of calculating induced velocities based on the values of 
circulation F, assuming them to be prescribed [see, for exanple, Baskin and Shi- 
Tstin (Ref.l6, 22)]. Therefore, it often happens that only the operation of de- 
termining the induced velocities with respect to prescribed values of F is in- 
troduced into the vortex theory concept. 

10. Constancy of Circulation of Trailing Vortices Along 
Straight lines Parallel to the Axis of the Inclined 
Vortex Cylinder and Possible Simplifications 

It was already noted in Subsection 2 that the circulation of spiral vor- 
tices is variable over their length because they merge with the radial vortices. 
Therefore, when calculating the integral in eq.(5.18T we must find the dependence 

^— = f(s?). In like manner, when calculating the integral in eq.(5.23) it must 



c^F 
be borne in mind that — ^— varies over the" length of the vortex sheet along 

with the variable «?. This circumstance coirplicates calculation of the integrals 
in eqs.(5.18) and (5.23). Therefore, in calculating these it is convenient to 

SF ?iT 
make use of the fact that s ^ and — -^ are constant along straight lines 

op ^tf 

parallel to the generatrix of the vortex cylinder. 

Actually, in the case of steady flow past a rotor, vortices of identical 
strength will be shed from a certain radius of each blade at azimuth f . These 
vortices will be carried away from the rotor along a straight line parallel to 
the axis of the vortex column. Therefore, at any distance from the rotor, at a 
point of the vortex sheet with azimuth j? = \|f and radius p = r, the strength of 
the spiral and radial vortices is identical. 

For further conputation, it is important to note that any straight line 
passing within the vortex column and parallel to its axis intersects the vortex 
sheet at the points 

where 
234 



n = 0, 1, 2, ...,<»; 

J? = azimuth of the rotor reckoned only in the range from to 2n; 
\|;n = azimuth of the blade for which the difference {^^ - ■&) has the smallest 
positive value. 

It follows from the structixre of eqs.(5.18) and (5 '23) that, to determine /221 
the induced velocity at some point of space, it is first necessary to integrate, 
over the entire vortex sheet, a function representing the product of the strength 
of an element of this vortex sheet and the induction coefficients L^ and M^,, and 
then to sum the resiilts obtained from the vortex sheet of each rotor blade 
separately. 

However, in this case we need not integrate along the vortices conprising 
the vortex sheet. At first, we can sum the products of the strength of an ele- 
ment of the vortex sheet and the induction coefficients L,^ and M^ along straight 
lines parallel to the axis of the vortex column. Here, by virtue of the 
strength of the vortices along the straight lines being constant, this operation 
reduces to summation of only the induction coefficients. Therefore, the ele- 
mentary conponents of the induced velocity from these vortices can be repre- 
sented as 



dvs 



dr^ 



P ="^ 2 ^N\dQM, 



dv 



dn 



rad=-^ '^M,^\dQd^, 



(5.31) 



where T^ is the circulation of the bound vortex at the instant when the blade is 
at azimuth f = ??. 

After integrating these expressions over the entire rotor disk, we obtain 
formulas for determining the axial conponents of the induced velocities in the 
foi-m 



\ n=0 / 

\ n=0 / 



rfO. 



(5.32) 



On the basis of these formulas, we can construct a conputational method 
applicable in practice. This method was first used by M.N.Tishchenko. 

It should be pointed out that, in the practical application of this method, 
the volume of conputational operations is very large. 

Thus, if for calculating the integrals in eq.(5.32) the circianference of 

235 



the rotor is divided vdth respect to azimuth into za sections and the blade into 
Zr sections with respect to radius, then for calculating the field of the axial 
induced velocity conponents in the rotor plane alone, the integrands in 
eqs.( 5 •32) must be calculated (z^Zj.)^ times. 

If we assume z^ = 72 (Aijr = 5°) and z^ = 30, then the quantity (zij,Zr)^ will 

be equal to about 4«5 ^ 10^. Therefore, this method can be effectively used 
only on conputers with a speed substantially greater than 20,000 operations per 
second. 

11. Characteristics pf^ Using the Iiftin^ Ij.ne Scheme 
and Scheme of a Vortex lifting Surface 

It was pointed out above that the scheme of a vortex lifting line yields 
satisfactory results if the induced velocities are calculated at a sufficient 
distance from the blade. However, for determining the aerodynamic loads by 
eq.(5.8) it is necessary to calculate the induced velocities on the blade, /222 
i.e., where the bouxid vortex is located according to the calculation scheme. 

If the lifting-line scheme is used, then the induced velocities in the 
calculations begin to increase on approach to the lifting vortex and vanish at 
the vortex itself. This takes place in two cases: 

1. If flow past the blade is unsteady and if radial (transverse) vor- 
tices are formed in the vortex wake. 

2. If the spiral (longitudinal) vortices shed from the blade make an 
angle differing from tt/2 with the blade axis, which always takes 
place in the case of oblique flow through the rotor since the blades 
have slip flow. 

Consequently, the induced velocities in the lifting vortex will not vanish 
only in the case of axial flow past the rotor. 

These difficulties can be overcome by neglecting the effect of radial voiv 
tices and rotor slip. Such an approach is widespread in practice and can be 
used whenever peraiissible with respect to the nature of the problem to be solved. 
However, this renders the solution rather approximate, which does not always 
siu-t the researcher. 

The method of calculating with the scheme of a vortex lifting surface is 
free from this shortcoming. Therefore, when calculating the induced velocities 
in the blade region we can use methods based on replacement of the blade ty a 
vortex surface, as is done in the theory of unsteady flow past an airfoil 
(Ref.30). However, this renders the problem of determining the induced veloci- 
ties even more conplex. Therefore, this approach is not yet in widespread 
practical lose for rotor calculation, although work in this direction is in 
progress (Ref.l9). 

For practical purposes, we can use the method in which the free vortices 
shed by the blade are divided into vortices directly adjacent to the blade and 
vortices remote from the blade. After this, the induced velocities due to 

236 



vortices remote from the blade can be determined by the vortex lifting line 
scheme while, for calculating the velocities due to vortices adjacent to the 
blade, a method based on the vortex lifting surface scheme can be developed. 

12. Division of Vortices into Types Close to and Remote 
from the Blade; Use of "Steady-Flow Hypothesis" 

To facilitate an analysis of the influence of various elements of the 
vortex sheet on the magnitude of the aerodynamic load on the blade, it is con- 
venient to divide free vortices into two classes (Ref.l7). The first class in- 
cludes vortices immediately adjoining the trailing edge of the blade in question 
and those shed from the blade during one revolution at some small azimuth angle 
Aijt (Ai|; = 20 - 30°). Such vortices are called adjacent (to a given blade) and 
the induced velocities caused by these vortices are called intrinsic. The 
second class includes all other free vortices. These vortices are called re- 
mote, and the velocities induced by them are called extrinsic induced velocities . 

Such a division is based on the fact that the vortices shed from the blade 
have a noticeable influence on the aerodynamic load of the blade only while they 
are sufficiently close to it. Upon removing the vortices by a distance of 
20 - 30° with respect to the rotor azimuth, their influence decreases but re- 
sumes its former extent when the blade executes one conplete revolution and again 
approaches these vortices. Thus, the blade along its path encounters not only 
its own vortices but also the vortices shed from all other blades of the rotor. 
All these vortices usually fall into the general group of remote vortices. /223 

In some cases, it is convenient in calculations to determine separately the 
velocity field induced by the remote vortices and to investigate the flow past a 
blade moving in this nonuniform field. With this approach, the flow past the 
blade will be similar to a flow past the wing of an airplane flying under condi- 
tions of t\u?bulent air. In the same manner as for an airplane wing, when calcu- 
lating the variable aerodynamic loads on a given blade, it is possible to con- 
sider the effect of trailing vortices directly adjacent to the blade by the so- 
called "steady-flow hypothesis". In this hypothesis, it is assumed that, in un- 
steady flow past a profile, the loads acting on the profile behave as though the 
flow pattern produced at a given instant of time remained imchanged for an arbi- 
trarily long time. In the "steady-flow l::ypothesis", the effect of trailing vor- 
tices adjacent to the blade is disregarded. 

Thus, the flow pattern of the blade can be represented in the following 
form: When the rotor rotates, the blade encotinters the nonuniform field of ex- 
trinsic induced velocities caused by the effect of the total vortex system of 
the rotor with the exception of the vortices immediately adjoining the blade. 
Under the effect of this velocity field, the angles of attack of the blade sec- 
tions vary constantly, and variable aerodynamic loads caused by the nonunifOrmity 
of this field begin to act on the blade. The magnitude of the variable aerody- 
namic loads is affected also by the free vortices adjacent to the blade and shed 
from it vipon any change in the circulation flow. The effect of adjacent vor- 
tices is of the same nature as that in unsteady flow past a finite airplane wing. 



237 



13. Instantaneous and Mean Induced yelocities and Ge neration 
of Variable Aerodynanac Xoads on the Blade 

Calculation of induced velocities by the theoretical scheme with a finite 
number of blades makes it possible to determine the true (instantaneous) induced 
velocity at any point of space near the helicopter. The true induced velocities 
prove to be variable, fluctuating in time with the frequency of the vortices 
passing by a given point. One can distinguish time-invariant (mean) and time- 
variant conponents of induced velocities. 

If we examine a point in space fixed with respect to the helicopter close 
to its rotor, then the time-variant induced velocities at this point will be 
caused both by bound and free vortices . However, in conparing their values, it 
becomes obvious that the variable induced velocities due to bound vortices and 
to free vortices adjacent to the blade are largest in value. These induced ve- 
locities fluctuate with the frequency of the rotor blades passing by the point 
in question. Generation of these velocities is related primarily with the forma- 
tion of aerodynamic forces on the blade and is observed when the load on the 
blade remains constant in time, for exanple, in hovering flight. The velocity 
field induced by these vortices in this case rotates together with the blade. 

Something similar happens in the case of oblique flow past the rotor in 
forward flight. Here, we can distinguish a certain conponent of the velocity 
field which, rotating together with the blade, does not excite variable aerody- 
namic loads on it. To calculate these loads, it is necessary to define the mode 
of variation of the induced velocities at a point rotating together with the /22U 
blade rather than at a fixed point of the rotor disk area. 

With this approach, the main generator of variable aerodynamic forces on 
the blade is the nonuniformity of the extrinsic induced velocity field caused 
by vortices remote from the blade. Therefore, in the first approximation we 
can neglect the velocity field caused by vortices directly adjacent to the blade 
and investigate only the field of extrinsic induced velocities. 

14. Characteristics of the Extrin sic Induced Velocity Field 

An investigation of the extrinsic induced velocity field caused only by 
vortices remote from the blade permits the statement that their variable portion, 
at a point in space fixed with respect to the helicopter, will be smaller the 
greater the density of these vortices. 

An increase in vortex density takes place, in particular, on any reduction 
in forward flying speed of the helicopter. The density also increases with an 
increase in rotor angle of attack, when the mean velocity of flow through the 
rotor decreases and the vortex sheet is not carried away rapidly enough from the 
plane swept by the blades. Such a situation specifically occurs in braking 
regimes of the helicopter before passing to the hovering state. As an exanple. 
Fig. 2. 131 gives a plan view of the vortex system shed only from the blade tips 
in a flying regime with ij, = O.O5. The pattern shown is inconplete, since only 
the three spiral vortices shed from the blade tips are given while the vortices 
shed from all other blade radii are omitted. The radial vortices are also left 

2?>& 



Direction of flight 




Pig. 2.131 View of a Vortex System 
Shed by the Blade Tip in the |j, = 
= 0.05 Regime. 



out . However, even this pattern gives 
an idea of the extremely close spacing 
of vortices in low-speed regimes. 

The variable portion of the ex- 
trinsic induced velocities decreases 
also with an increase in the number of 
rotor blades. At the limit for a rotor 
with an infinite number of blades, the 
variable conponents of the induced ve- 
locity become equal to zero. 

To calculate the extrinsic induced 
velocity field, it is possible to use a 
scheme of a rotor with an infinite numn 
ber of infinitely narrow blades. This 
scheme jd-elds more accurate results, /225 
the greater the density of the free 
vortex system of the rotor in the 
flight regime in question. 



Upon changing from a given rotor 
to a design with an infinite number of blades, the local effect due to vortices 
innnediately adjacent to the blade is reduced so greatly that, in first approxi- 
mation, it can be assumed that this scheme does not allow for the effect of ad- 
jacent vortices, so that the field determined from this scheme will be closer to 
the extrinsic induced velocity field, the greater the density of the free vor- 
tices. Thus, a direct application of this scheme to the determination of vari- 
able aerodynamic loads on a given blade is equivalent to the use of the "steady- 
flow hypothesis". 

The scheme with an infinite number of blades can also be used whenever it 
suffices to determine only the time-average part of the induced velocity. Use 
of this scheme leads to substantial sinplifications of the problem and eliminates 
many difficulties arising with the scheme of a finite number of blades. Spe- 
cifically, one of the advantages of this scheme is the fact that the induced ve- 
locities nowhere vanish. 

In detennining the downwash at the wing and stabilizer of a helicopter, 
when we are usually interested only in the constant portion of the loads, the 
scheme with an infinite niMber of blades can be used in all flying regimes and 
yields conpletely satisfactory results. The same probably holds for determina- 
tion of the mutual interference of rotors, if the designer is interested only in 
their integral characteristics. 

15. Vortex Theory of a Rotor with an Infinite Number of Blades 

The vortex theory of a rotor with an infinite number of blades has been 
quite thoroughly developed. The solution of this problem was discussed specifi- 
cally in the works of G.I.Maykopar, A.I.Slutskiy, L.S.Vil'dgrube, A.M.Proskur- 
yakov, V.E.Baskin, Wang Shi-Tsun, and other authors. Each of them brought this 



239 



theory to e-vBr gireater perfection. 

While working o\it the vortex theoary of a rotor- with an infinite number of 
blades, mar^ methods were suggested ■vdiioh enployed certain additional assunp- 
tions-: 

1. The vortex sheet trailing from a rotor is plane-. This assunption formed 
the basis of the works by L.S^Vil*dgrube and L.O-.Mel»ts-, ard substantially 
sinplifies the calculations.. Therefore., such an approach, wJiLoh L.S,.Fil*dgrube 
brought to a form convenient for practical use, became widespread-. 

2. It was proposed in a number of p^ers that the irduced velocities can be 
determined rather accurately, with consideration of only the constant portion 
of the circulation of bound vortices or with ttre addition of one or two first 
haimonios of this circulation. 

Recently,, V*K.Baskin (Eef^l6'.) and Wang Shi-Tsun ('Itef-.22) published papei^ 
in which they discarded these additional assunptions and brought the method of 
such calculation to a state conpletely convB^nieiirt for practical use- Therefore, 
an accotmt of only these two methods will be given below. 

Vortex Theory of Wang Shi--Tsun_ 

The vortex theory of a rotor proposed by Wang Shi-Tsun has been ratter 
thoroughly presented in the anxthoi^s work (Ref-SZ^^/. Therefore, only the basic 
results will be repeated here- 

16. Rotor Scheme 

In all vortex theories examining a rotor with an infinite number a£ bUadea,- 
it is proposed that the surface swept by the rotor is covered by continuously /22& 
distributed 2?aidial bound vortices with circulation vardablLe ovra? its radius and 
azimuth. The surtfaee swept by the rotor is corsidered to be plane- The cord-ng 
angle of the blades is disregarded- 

Circulation of the Dound vortices located aloing the radius in the rotor 
sector with an angle A:i?-is taken as equal to 

where 

r = circ^alation in the blade section at the examined radius and azimuth 

of the rotor; 
Zt = nimiber of blades in the i*otor- 

It is further assumed that t-here exists a spiral (lon^tudinal'i) free vortex 
with a ciroiLlation of 

d^p )-^ '^d^^» (5.33) 

240 



w 



due to the rotor element imder study. 

As a consequence of the circulation of the bound vortex, varying with any 
change in the azimuth position of the blade, radial (transverse) vortices vjith 
a circulation of 

are also shed by the element in question. 

The form of the sTorface over which the free vortices are arranged is repre- 
sented as a downwash spiral surface, jxist as in the vortex theory of a rotor 
with a finite number of blades . However, now the vortex cylinder is filled with 
continuous free vortices rather than with discrete layers of widely spaced vor- 
tices j as had been the case in the theory with a finite number of blades. 

17. D etermination of Induced Velocities 

To find the induced velocities due to trailing and bound vortices we used 
the Biot-Savart formula [eq.(5.6)]. As noted above, to determine the total in- 
duced velocities, it is necessary to sian all elementary induced velocities ob- 
tained from individiml elements of all vortices conprising the vortex system. 
For this^ integration in the form of eq.(5.7) must be carried out. Wang Shi- 
Tsim demonstrated that there is no need here to integrate along the downwash 
spiral lines along wMoh lie the trailing vortices shed by the blade. It is 
sinpler to integrate along the straight lines parallel to the axis of the in- 
clined vortex cylinder (AB in Fig. 2. 129), since the vortices have identical 
strength along these straight Unes. Such a method was used earlier by I.O. 
Mel*ts for the case of a pla,ne vortex system. 

Thus_, in calculating induced velocities for the scheme of a rotor with an 
infinite number of blades, it is possible to carry out integration along straight 
lines parallel to the axis of the vortex cylinder rather than summation of indi- 
vidual discrete quantities, as is the case in the scheme with a finite niimber of 
blades (see Subsect.lO). 

Using this fact, Wang Shi-Tsun was able to calculate the above integrals 
and to obtain sufficiently sinple formulas for detennining all induced -velocity 
conponents.. 

IS. Calculation Formulas for Induced Velocity Determination 

The induced velocities (Ref..22) are i^resented as the sum of three types: 

■"^■Vbo +^sp +'"rad' (5.35) 

where _ /227 

V = total induced velocity; 
v^,, = induced velocity due to bound vortices; 

241 



_vbp 



induced velocity due to spiral (longitudinal) free vortices; 



Vpad ~ induced velocity due to radial (transverse) free vortices. 

All induced velocity conponents, entering eq.(5.35)» pertain to the peri- 
pheral speed of the blade tip cuR. 

Of greatest interest for practical application are the axial conponents 
(parallel to the rotor axis) of the induced velocity. In setting \sp the problem 
of calculating the rotor blade loads, axial induced velocity conponents need be 
determined only in its plane. 

We will give the calculation fonnulas for determining the axial conponents 
in the rotor plane only. The formulas for other conponents, determinable out- 
side this plane, can be found elsewhere (Ref.22). 

The axial induced velocity conponents due to bound vortices are determined 
by the fonnula 



I 2it _ 




(5.36) 



where T is the circulation in the blade section referred to cuR^ : 






(5.37) 



The axial induced velocity conponents due to spiral and radial vortices are 
determined by the formiilas 



Here, 



where 



V. 



'sp 



__ffc_ 






X 



8n2 Y^ + x= 

12b _ 

X j" r|£f(|x-f esin&yi-Qcos&/2ld»a:e, 







12r. 



■"r«^= 



_ ^b 



8ll2 



/(^^ + ^L 



X 



m^ 



cos &/j -]-sin ^/j] d^dq. 



/,- 



(C sin 0— r sin <1/) j/^|ji2 + X 



2 
oav 



^2 = 



/2 |/(l2 + ^oav +f^' (e COS » - r COS i/) 
(- Q COS » + Fcos ^) ]/"fl^ + X^^^-Tft 



/2]/ ^2 + X2^+ p./ [q cos » - r COS ^ ] ' 

/"= Yq"^ +^^ — 2q r cos(& — <jO. 



(5.38) 



(5.39) 
(5.40) 



242 



For the variables entering eqs.(5»36) and (.5'3S) we will use the following 
notations : 

p" = relative radius of the blade section shedding the vortex; 

r = relative radius of the rotor at which the induced velocity is de- /22S 

termined; 
2? = azimuth of the rotor shedding the vortex; 
i|r = azimuth at which the induced velocity is detemiined. 

Equations (.5'3^)f (5»36), i5'3&), and (5*39) permit determining the induced 
velocities if the circulation T(^, p ) is known at all azimuths «? and radii of 

the blade. 

n 



All 



% 



i 



■-^' 2 3 



^ 



8 9 



% 



%\ 



10 II 



Fig. 2. 132 Distribution of Circula- 
tion over Blade Length, Used in the 
Calculation. 



It should be noted that, in the 
vortex theory of a rotor with an infinite 
number of blades, one usually determines 
the total induced velocities, including 
the conponents due to bound vortices 
participating in the production of blade 
lift. 



19. Applicat ion and E valiiation of the 
Possibilities of the Wang Shi-Tsun 
Vortex Theory 



In the practical application of the 
Wang Shi-Tsun vortex theoiy, just as in 
a number of other schemes, the distribu- 
tion of circulation V over the blade length is usually represented as a stepped 

line, as shown in Fig. 2- 132. The quantity ^_ ' dp is taken as approximately 
equal to 



Bp 



dr 



dQ-=r,-r,_^^i:,rQ, 



{5. hi) 



where AF^ is the difference of circulation at two adjacent portions of the blade. 

Such an approach is equivalent to replacing the vortex sheet by a number of 
discrete vortices. As a consequence, the induced velocity approaches infinity 
at points where these discrete vortices are shed by the blade. To avoid this, 
the induced velocities should be calculated with respect to the midsection at 
constant circulation. 



In determining the circulation derivative with respect to blade azimuth, 
the circumference of the rotor is also divided into a finite number of sections 



and the derivative 






da? is taken as approximately equal to 



^d^-r,-^,_,^^^,: 



<.5.k2) 



2k3 



Thus, for a nimierical determination of the integrals entering eqs.(5«36) 
and {5'3s)f we should first determine the values of Fq, AF^, and AF^ , whose num.- 



ber N -will be equal to 



where 



//=: 






(5.43) 



Zr = ntmber of sections over the radius into which the blade is divided; 
Aijf = pitch with respect to blade azimuth over whose length the circula- 
tion is considered constant. 

If we take z = 30 and A^f = 5°» a determination of the induced velocities 
will require 2l60 values of Fp, AFp, AF^ (a total of 64S0 values). To determine 
the induced velocities at all 2l60 points of the rotor, the operations needed /229 
for determining the integrands in eqs.(5.36), and (5.38) would have to be re- 
peated (2160)^ tim.es. It is obvious that the time required for this would handi- 
cap the practical applicability of this method, even when using high-speed com- 
puters. Therefore, the practical application of the Wang Shi-Tsun theory in 

such a form is possible only for a 
rather limited number of sections, 
taken over the radius and azimuth of 
the blade. On high-speed conputers 
with a rate of 20,000 operations per 
second, the calculation can be per- 
formed under approximately the follow- 




ing conditions: 
= 10 - 12°. 



Zr = 10 - 12; At = 



Fig. 2. 133 Diagram of Formation of 
Vortex Rings. 



This shortcoming is eliminated if 
the circulation F and induced veloci- 
ties are e^qjanded in Fourier series in 
harmonics and if the process is limited 
to considering only a certain number 
of lower harmonics necessary for per- 
fonning the calciilation with satis- 
factory accuracy. For practical pur- 
poses, it usually suffices to restrict 
the process to the first 6-8 har- 
monics. This method was used in the 
theory of V.E.Baskin (Ref.l6). 



Vortex Theory, of V.E.Baski.n 

The theory of V.E.Baskin was circiolated in a limited edition in 1955 and 
later presented by the author in an urpublished report (Ref .16); as a conse- 
quence, the theory of V.E.Baskin is not very familiar to specialists interested 
in this problem. For this reason, the theory will be presented here in consider- 
able detail but with certain sinplifications which the author did not make. 



2hk 



2Q. Sc heme of Rotor Elxgw 

The vortex system, shown in Big^2.'l2S and. consisting of bound and free 
spiral and radial vortices, can be represented in a somewhat different form. 
We can consider that, during operation of the rotor, the blade continuously 
sheds infinitely small closed elementary vortex rings, with a circulation con- 
stant over the contour of the ring etjial to the circulation of the bound vortex 
r in the blade section at the instant at which the vortex ring separates from 
it ('Fig.^2..13S0-. 

Since the circtilation of the bound vortex varies over the radius and azi- 
muth of the blade, the vortex riiags trailing frean it will have a different 
circulation at different points of the vortex sheet. Consequently, . the free 
vortex sheet can be considered as consisting of continuously distributed, in- 
finitely small vortex ri-ngs with different circulation T . 

It is known from hydrodynamics Usee (iRef.^l, p..-266.).'] that the induced ve- 
locities due to the vortex ring contour with a 
^i circulation T will be the same as those due to 

the layer of dipoles covering the surface 
stretched over this contour, with an intensity 
<ojQ^ per unit area of 



230 




B — r 



(5.44) 



Fig.-2.134 For Determining 
Induced Velocities from a 
Dipole Goliann. 



and oriented along a normal to this surface. 
Here, the signs are selected on the basis that 
the circulation flow producing blade lift and 
the dipoles whose vector gives a positive projec- 
tion onto the y— axis, are considered positive 
(iFig.-2.134X. 

Consequently, the free vortex sheet trailing 
from the blades can be replaced by a surface 
covered by a dipole layer. 



On changing to a rotor with an infinite 
number of blades, this equality is somewhat modi- 
fied. Aft of a rotor with an infinite number of blades, the vortex sheet fills 
the entire volume bounded by the downwash cylindrical surface tangent to the 
rotor circumference. This vortex system, as already mentioned, is called a vor- 
tex column. This column can be represented as fill fled with dipoles rather than 
with vojrtices. To determine the intensity of dipole distribution in the col\min, 
a layer of height dH, filled with vortices during the time dt, is cut from the 
coltimn. During this time, vortex rings with a contour boimding the area 



dF==UattiQ 



(5.45) 



trail from all radia_. 

This contour can be replaced by dipoles with an intensity D = -F . On 
changing to an infinite number of blades, these dipoles are distributed over the 



245 



entire rotor circimiference of radius r with an intensity of 

^=-|7^''^- (5.46) 

If the circulation T is ej^jressed in terms of the aerodynamic load per unit 
length from eq.(5«24), then 

D=-£iI^. (5.47) 

2iter 

We can determine the quantity dt entering this expression as 

UN 



dt=- 



(5.48) 



where Vy is the average velocity of flow through the rotor: 

Let us then introduce the concept of relative aerodynamic load, so that /231 

T=^ciebo,^WP. (5.49) 



Then, 



4 " '■ ^y 



Thus, the problem reduces to determining the induced velocity field caused 
by the downwash column of dipoles whose intensity in a layer of thickness dH is 
determined by eq.(5.50). 

21. Determination of Induced Velocities from 
the Dipole Column 

Figure 2.134 shows that the plane of each of the elementary vortex rings 
trailing from the blade is inclined at some angle i to the plane of rotation of 
the rotor: 

o«^. (5.51) 

Correspondingly, at this same angle to the plane of rotation are inclined 
also the axes of the dipoles which, in V.E.Baskln's scheme, replace the ele- 
mentary vortex rings. 

Henceforth we will disregard this angle and will assume that the axes of 
all dipoles are directed perpendiciilar to the rotor plane. Baskin (Ref .16) did 
not use this slnplificatlon in deriving the basic relationships, but it is 
shown that - in calculating the axial induced velocity conponent - this is per- 
missible . 

246 



To determine the induced velocities, let us examine a flow caused by di- 
poles filling an element of the vortex cylinder of a height equal to dH (see 
Fig. 2.136). 

Summing the induced velocities caused by all elements of the vortex column, 
we can obtain the total induced velocity. 

22. Fluid Flow Induced bv a Disk Covered with Dipoles 

Fluid flow induced by a disk covered with dipoles can be determined as the 
solution of the laplace equation with certain boundary conditions, which will 

be discussed below. 



u . 



The laplace equation in cylindrical coordinates 
(Fig. 2. 135) can be written in the form 






duZ "■" (Jq2 ""^ e dQ q2 (JS2 ~" ' 



(5.52) 



Fig. 2. 135 For De- 
tenninlng the Ve- 
locity Potential by 
the laplace Equation. 



where 9 is the velocity potential. 

The velocity potential can be ejqsanded in a 
Fourier series. Then, 



= 2 ('PmCos/ra»+ cp„SinOT»). 



(5.53) 



m=0 



Substituting eq.(5.53) into eq.(5.52), we obtain the equation for determining 
the coefficients of the sines and cosines entering eq.(5.53)! 






Solution of eq.(5.54) will be sought in the form 

9« = e^*''5C(0). 
Substituting eq.(5.55) Into eq.(5.54), we obtain 



x"(Q)+-^x'(G)+(^^-^)z(0)=0. 



(5.54) 

/232 
(5.55) 



(5.56) 



This equation can be reduced to a canonical form, if we set 

Aq=q. 

Then, 



247 



The general solution of this equation can be .vretitten in the fonn 

The ftmction I^Cp^) tends to infinity as p" -» Q. Therefore, if we restrict 
ourselves to a finite value of the potential, then 

<P,„=^e±*".y„(*Q). (5,.59) 

If we use the condition that, as u -♦ <= also cpo " 0» then the solution of 
eq.(5«54) can be written as 

<P,„=yie-*«y„(,*Q), (5.60) 

where k 2: Q. 

The function cp„ will be the solution to eq.(5-.5'^) at any value of k = kj,. 
Consequently, the solution of eq.(5-«54) will also be the sum 

<P.„.=2^i«~*'"^(^v.Q) ^5-61) 

and the integral 

'P.nr^JA{k)e-'»'J„(jkQ)dft, ('5.62) 



where A(k;) is an arbitrary function detennined from the boiindary conditiona. 

23.. Bourdaaw Gonditions 

It is known from hydrodynafflics tsee (iEef-.-31, p..-82)j] that the velocity po- 
tential on both sides of a surfaoe covered ty dipoles is associated with the 
density of these dipoles by the relation 

il-=<P2'— <pi, (5-. 65) 

vheve 

D = density of dipoles per unit surface, with the positive direc- 
tion of the dipole asd-s coinciding with the u-axis-; 
91 andcpg = velocity potentials on both sides of a surface covered by 
dipoles (iEig.-2.136:)-. 

The fluid flow on both sides of the disk covered by dipoles is subject to 
equal conditiona. Therefore, we can set 

fjt=— <pi=(p. (5.64) 

248 




Fig. 2. 136 Diagram of Rotor Flow Used in 
the Calculation for Determination of 
Induced Velocities. 



Then the value of the po- 
tential at the boundary of the 
disk on the side where u > 0, can 
be defined as 



9=-fD. 



(5.65) 



The boiindaiy conditions for /233 
flow induced by a disk covered by 
dipoles of a density D are deter- 
mined from eq.(5.65) if p ^ R. If 
p > R, the potential in the disk 
plane is everywhere equal to zero 
(cp = 0),. 

The intensity of the dipoles 
covering the disk can be repre- 
sented in form of a serieg:: 



-^j9rt(B)sIn/rt6]. 



('5-.^6.) 



Then, using the boundary condition ('.5-.65-) for detewQining the ai*bitrary 
function A(k) in eq.(5«62), we can write the Solution of the (jontiniiity equa- 
tion (5 '52) in the form 



9=-- ^2]J'^''*"-^'«(*e)*x 

m=0 

The derivation of this ejqjression is given elsewhere (Ref.3l)-. 



X 



(5.67) 



24. Transform ation of Eg. (5.6?) t o th e Rotor Axes; Use of 
the Theo rem of Addi tion of _gylindricai Functions 

Equation (5.6?) is written in cylindrical coordinates with the u-axLs going 
through the center of the disk of thickness dH, cut out from the dipole column 
(see Fig. 2. 136). To transfonn this ejqpression into coordinates related with the 
rotor, we can use the theorem of addition of cylindrical functions (Ref .34). /234 
It follows from this theorem that 



cos/n»y„(Ac)=(-l)« 2 Jnikr)Jn-^m{kl)cosn<^, I 



(5.68) 

249 



sinm^J„(kQ)=-{-^iy" 2 J„ikr)J„+„{kl)sinn'^. 



The velocity potential from the dipole layer is ejipanded in a Fourier 
series: 



?= '^i'?nCOsn<^-\-<f„sinn<^). 



(5.69) 



Then, using eqs.(5«68) and (5 '69) and equating the coefficients of cos nj(f 
and sin raji with identical n in eq.(5.67), we obtain 



<?«= 



■2( _ 1 )m j e-*„ y^ (^kr) y„^„ (kl) kdkX 

m—O 

XJ5„(e)-/m(*Q)Q"fQ, 

) 

^(-ir ^ e-'"'J„(kr)J„+Jkl)kdkX 

m=0 

R ^ 
X^D^(Q)J„(kQ)QdQ. 





IS 



(5.70) 



25. Determination of the Total Velocity Potential from 
the Entire Dipole Column 

To determine the conponents of the total velocity potential represented as 

^= '^(^nf^osn.'^ + ^^sinn'^), (5.71) 

eqs.(5.70) must be integrated over the entire dipole column (see Fig. 2. 136). 

For this, we must first write out the values of D„(p) and D„(p). 

We note that within the vortex column the intensity of the dipoles, just as 
the strength of the vortices, is constant along straight lines parallel to the 
column axis. Therefore, the dipole intensity D(p , ^) in arQr layer cut out from 
the vortex colimm will be equal to the intensity of the dipoles trailing from 
the rotor at a point where r = p and ^ = ■&. 



If the relative aerodynamic load P [see eq.(5.50)] is represented in the /235 



form 






(5.72) 



250 



then we can write 

4 " Vy Q 

4 " Vy Q 

Setting 

^=^7^1 (5.74) 

and substituting eqs.(5»73) into (5.70), we can integrate the obtained expres- 
sions with respect to H. It is easy to demonstrate that, for perforaiing this 
operation, we must determine the value of the integral 

JiH) = °°^e-''^J„^„(^k'^H)dH. (5.75) 

Referring to the handbook by Ruzhik [Ref .34, p. 721, eq. (6. 611.1)], we obtain 

JiH)=±^A,TP, (5.76) 

where 

"""^yTT^: (5.77) 



■"- ,. ■• (5.78) 

Here p = m + n, and eq.(5.76) is valid only if p > 0. Therefore, at m + 
+ n < 0, eq.(5.76) takes the form 

M/^)~^A,{~\y'cP, (5.79) 

where p = -(m + n). 

Let us introduce the new variable 

z=kR. (5.80) 

Then, using eqs.(5.76) and (5. 80), we can write out the expressions for /236 
the conponents of the velocity potential from the entire dlpole column: 



251 



X ] e-^~"J,{z?)dz \P„(Q)J„izQ) dQ, 

m=o 

X J e-^~"J,iz7) dzlp„{Q)J^(zQ)dQ. 



(5.81) 



Here, all linear dimensions pertain to the radius of the rotor R, including 



y(? = ^). 



26. Determination of^ Inducted Velocitie s 

As already mentioned, we will determine only the ajcLal induced velocity 
conponents. For tMs., we must take the derivative of the velocity potential 

with respect to y. We see from eq.(5.81,) that only the term e~^^ depends on y,. 
Therefore, the operation of differentiation leads to e:^ressions differing from 
eq.(5«81) only in sign and in conppnent z. 

Before writing out the final formulas for determining the induced veloci- 
ties, we will present them as a Fourier series: 



^= 2'(^n cos «++X„ sin «<];). 



(5.82) 



Here, all induced velocity conponents are referred to the tip speed of the 
blade (uR. In determining the induced velocity conponents written in form of 
eq.(5.82), the following operations must be performed: 



" <^R\dy dy )' 

)• 



dy 
" u/? V dy 



dy 



{5'&3) 



An exception is the determination of Xq, which is calculated as 

r- 1 _ ^«g"o 

0)/? dy 



(5.84) 



As a result of the actions provided for in eqs.(5.83) and (5.84), we obtain 



252 



o' 



M 

m=0 

X ] e-'~«j„{£'r)zdz J ^„(g)y^(zo)de. 



Z222 



(5.85) 



('5-86) 



It is assijmed in these expressions that n > m. Therefore, in conformity 
with eq.(5.79)^ when n < m, in place of t"""" we must calculate (_ii)°'-°j»-° . 

When performing operations with eqs-.(5-«83) we must bear in mind that 

y_„(£:7)=(-i)''y„(z7.). 

Different methods can be s'uggested for calculating the irttegrals entering 
eqs.(5»85) and (5^86)^ One such method will be given in Section 8 of Chapter I 
of Vol.. II, where a method of calculating elastic vi'brations of a blade with con- 
sideration of a variable induced velocity field will be disxjusised-. 

It should be pointed out that, in certain particular cases, the integrals 
eriterirg eqs-.('5-.85^ aiid ('5-* 86) can be calculated analytically (see Sect -8-. 6 of 
Ohapt-I in Vol- II) ^ 

The vortex theory af V-E-Baslcin permits obtaining results rather rapidly, 
if the person doing the calculations has at his disposal a ccoputer with a rafte 
of at least 20,000 qperations p&r second- At this rate, the calculation cain be 
penformed in about ^ min. Therefore, if the problem is to determine the mean 
induced velocities with an accuracy to high harmonics and if the assunption that 
the vortex sheerfc is plarne cannot be used, the best method for a determination Of 
induced velocities is the Baskin vortex theoiy.. 

Section 6- Bsgaerimental Determination of Aerodynamic 
Qharaeteristios of a Rotor 

The most reliable methods of determining the aerodynamic characteristics of 
a rotor are flight tests of the helicopter with the rotor under study or with 
another similar rotor wlrdeh can be regarded as a model of the investigated rotor, 
as well as wind-tunnel tests on full-scale rotors or large-scale models. 



253 



In this Section, we -will present certain results of flight tests and wind- 
tunnel tests and give a brief description of the testing procedure. The tests /238 
were carried out at research institutes ty M.K.Speranskiy, A.I.Akmov, and 
others . 

1. Flight Tests for Determining the Aerodynamic 
Characteristics of a Helicopter 

The aerodynamic characteristics of a helicopter in a system of wind axes 
(Figs. 2. 105 - 2.109), i.e., in the form in which they are used in aerodynamic 
design, can be obtained from flight tests. 

Selecting the flying speed, rotor rpm, flying weight, and flight altitude, 
flight tests will furnish constant values of M^j^ , Mq, and the thrust coefficient 
of the helicopter ty (below, p and T denote the pressure and tenperature of the 
air) : 

V 



f^fl = — = 7r= = const, — ?^; 

^ a 20.1 /r ' -/t ' 

Mo=- = const2-^^; 

a y T 

since ±- Qa= = -^ (o.379 ^) (2Q.1 Yff=com{,p. 

Fulfilling, under these conditions, various flight regimes with different 
engine powers - gliding at a different rate of descent, forward flight, and 
climbing with a different rate of ascent, we can obtain a sufficient number of 
points of the aerodynamic characteristics of the helicopter in the form of the 
dependence t^ = f (nit ) for constant values of V, ty^ , and Mq . In an autorotation 
regime of the rotor, we obtain a point with m^ « (Fig. 2. 137). The point t^ 

= corresponds to a forward flight regime since, at this point, the propulsive 
force of the rotor is balanced by the parasite drag of the helicopter. Gliding 
of the helicopter corresponds to intennediate regimes. The regimes t^j^ < 0, 

i.e., a gain in altitude at a point where the engine power is maximal (mt = 
= mt ), are obtained from the regime of maximum rate of ascent for a given 

flying speed. 

As a result of tests, we will determine the aerodynamic characteristics of 
a helicopter which differ from the rotor characteristics in that the aerodynamics 
of the no- lift-producing parts of the helicopter is taken into account. 

Thus, flight tests for determining the aerodynamic characteristics of a 
helicopter involve "flying by the seat of the pants" in which gliding, gain in 
altitude, and forward flight are performed at constant values of V//T, n//Y, -^. 

It 

254 





Hfi ''Const 


^ 


Autorotation VjCOHSt 




\ ty=const 


t 


\ Gl irfing 




\ „it"ei 




\ 1 flight 




\ ""i 




Gain in altitude—^ \ 



In addition to these quantities, the follow- 
ing are measured: vertical velocity Vy ; in- 
clination of the plane of rotation of the 
rotor to the horizon (pitch angle of the 
helicopter) ■&; conponents of cyclic change 
of pitch cpi, ^ (or angles of deflection of 
the automatic pitch control h, T] and the 
flapping coefficients); setting angle of the 
blade 9o ; and torque of the rotor M,. . The 
torque is measiored iDy strain gages mounted 
to the rotor shaft or to the rod of the re- 
duction-gear frame. 

The aerodynamic characteristics of a /239 
helicopter are determined from the e^^jres- 
sions derived in Section 1 of Chapter III: 



G cos 6ji.p 



1 1 



(6.1) 



Fig. 2. 137 Coefficients of 
Forces and Torque of Rotor in 
Various Flight Regimes. 



/«^ = 



Mf 



^= 



Q(u>R)2RaF 



G sin Qfi,p 



1 



-,Q((»R)2aF 



(6.2) 
(6.3) 



The flight-path angle to the horizontal is 



Ofifi = Si'n 



-/^v 



(6.4) 



-Equivalent plane 
of rotat ion 




The rotor angle of attack and 
the equivalent angle of attack are 
found from the following formulas 
(Fig. 2. 138): 

a^%-QfLp; (6.5) 

ae = a — (p,=a-|-£^i* — OjTi-l-Afe,, (6.6) 

where 

DiK - DgTl = deflection of the 
equivalent plane 
of rotation of 
the rotor from 
the design plane 
of rotation at inclined automatic pitch control; 
kbi = additional deflection of the equivalent plane of rotation of the rotor 



Fig. 2. 138 For Determining the Angle of 
Attack a and Equivalent Angle of Attack 
Q-g in Flight. 



255 



B 



in the presence of a flapping ccaipensator. 

At a methodologically correct conduction of tests, the root-mean-square 
error in determining ty is 0.5^ while, in determining mt, it is 3«5^« 

The results of flight tests with the helicqjter Mi-1 are plotted in 
Fig. 2. 139. The helicopter had a three-blade rotor with trapezoidal twisted 
blades, D = 1!(..3 m, a = 0.0504, blades with plj^ood planking except for the /240 
blade root (r < 0.59) where the shank portion of the profile was covered with 
fabric . 




-m -^ 



-0,02 



-0.03 



-OM 



III 1 ^ 1 L 


,ot=6,Z' ^Mfi =0.1 -- 


V =0.20 
("-^ M,=O.S ~^ 


^ 


St ^ 


a5 l^lt 


A^ 


XA 


S^ 


^^ 


-^^^ J^ 


^X 


^tt. 


X S^ 


^- -2-5°A^ 


U 


SA 


- -zr^^V/Z',- 


aoos'\\\ 0,010 ml 


-^WU-L. _j 


v\ > 


\\\h-0J7 


tr0.n\\ 


\W,^-rOJi 


V\'\ 1 


\V ^ fCI 


-75,-l:\a;^ 


15...-^, 


-I^S'W 


-+- '^-^-v^ 


:5tK 


^ 


± 



Fig. 2. 139 Coefficient of Prqpulsive Forces as a Function of 
Coefficient of Torque (Flight Tests of Mi-1 Helicopter). 



The aerodynamic characteristics of the rotor are obtained by subtracting 
the coefficients of forces created by the nonlifting elements from the coeffi- 
cients of forces of the helicopter. Ely analogy with eq.(l.3) from Chapter III, 
the coefficients of forces of the rotor are equal to 



256 



- v^ 

- V^ 
^x = ^x. - ^Xf.^r= ''A ~^'~' 



(6.7) 
(6.8) 



where c 



Cy^S, 



yf 



and Cj = 



Sc^S 



The coefficients Cy and c^ of the Mi-1 helicopter were determined hy test- 
ing a fiill-scale fuselage in a wind tunnel. The lift coefficient of the fuse- 
lage 'Cy is very small and we can disregard it in eq.(6.7). 

It is necessary to bear in mind that the aerodynamic characteristics of /2hl 
a rotor obtained from flight tests take into account the mutual interference of 
the lifting and nonlifting elements of the helicopter, a fact that increases the 
value of these data. 

2. Wind-T'unnel T ests for Determining the Aerodynamic 
Charac teristics of a Rotor 

To determine the aerodynamic characteristics of full-scale helicopter 
rotors in a wind tunnel, the TsAGI has special facilities for testing two-, 
three-, and four-blade rotors with a diameter vp to 15*5 m. 




Fig.2.15t.O Facility for Testing Full-Scale Rotors 
in a Wind Tunnel. 



The first facility for testing full-scale rotors with an engine power of 
575 ip was created imder the sipervision of M.L.Mil' on the basis of the Mi-1 
helicopter. Figure 2.1^0 gives a view of the Tiiiit mounted to the vpper structure 
of the tunnel balance. The unit has a special danping suspension support, main- 
taining the permissible level of vibrations set vp on the balance during start- 

257 



t 


1 e -»: ' 


020 






J!> 








_^ 




,^ 


Oil! 


^ 




,• ^ 


to 


> 


0-/ 




* 








VO g'i ^ 




7^ ^ 




7 




6f 








VJJO Of 


■ 


«i 




1 




ft';v 









£1005 



0.010 



.0,015 m^ 



Fig. 2. 141 Aerodynamic Characteristics 

of Rotor (V = 0; Mo = 0.5; o = 0.0525; 

Experiment) 



/gig 



Fig. 2. 142 Aerodynamic Characteristics 
of Rotor (Mfi = 0.075; V = 0.15; Mq = 
= 0.5; a = 0.0525; Experiment). 




-aoi 



258 




Fig. 143 Aerodynamic Characteristics of Rotor 

(Mfi = 0.1; V = 0.2; Mq = 0.5; o = 0.0525; 

Experiment) . 




Fig.2.1M- Aerodynamic Characteristics of Rotor 
(Mfi = 0.125; V = 0.25; Mo = 0.5; o = 0.0525; 
Experiment) . 



JO 



up, overspeedlng, and nonnal operation of the rotor. 

The aerodynamic forces and moments of the rotor in the axes of the wind 
tunnel are determined as the difference between the readings of the balance with 
the special unit operating, as well as with the unit minus rotor and hub (or with 
a nonrotating hub). The obtained moments and forces include the mutual inter- 
ference of the rotor with the fuselage of the unit and the effect of the rotor 
hub. In some cases, corrections are introduced which take into account the ef- 
fect of the rotor on the forces created hy the fuselage of the unit. In such 
cases, and also when measuring with a strain-gage balance placed near the hub, 
the characteristics of the rotor include only the effect of the fuselage on the 
rotor and the effect of the rotor hub. 

Figures 2.141 - 2.145 show the test-derived aerodynamic characteristics of 
the rotor of the Mi-1 helicopter, with metal blades of rectangtilar planform. 
The geometric characteristics of the blade are plotted in K.g.2.146: blade pro- 
file NACA 230, number of blades z^ = 3, diameter D = 14.5 m, solidity ratio /244 
CT = 0.0^25t inass characteristics of blade y = 4«5, flapping conpensator k = 0.56. 



tx 


^ 














— 


-- - 








— 


— 




\ 
















\ 


k 










ty'C 


.20 




^\^ 


^^ 


^ 


■^ 


K 




^^0.16 

^o.n 

,^0.12 
Xy=0.10 








~ 




sjv^ 


s> 


\^ 




^ 












*N^\ 


^\ 


y 




r-' 












0.01 


^ 




y 






















^ 






















V 






\ 










■ 


— 


- 




c^=-2*^^ 






s\ 


1 






- 






cC=-if' 




\\\ 


\ 







oC=-6°\ 


^. 


oos 


\ 


\ 




mi 




r 


"f 








1 


^ 


% 


v\ 


1 


1 


^ 


- 


~ 






— 


— 




^ 




7° 


% 


1 
















X 


1 


^ 


-0,01 
























OL=- 


-12°^ 














- 


'^ 


- 














\\V 


















_ 


V^ 


~0,02 


























' 
















^V 



I 


W 










-f 








, 


- 


















0.05 








\ 






N 


1 










i 








N 






&if> 










\ 














X 


N 
















N 


















5 




N 


. w 












- 


— 


N 


- 

























If 
z 

f 

-z 



Fig. 2.145 Aerodynamic Characteristics Fig. 2. 146 Geometric Characteristics 
of Rotor (Mfi = 0.15; V = 0.3% Mq = of Blade. 

= 0.5; a = 0.0525; Experiment). 

The escperiment was laid out so that the aerodynamic characteristics in- 
cluded the effect of the rotor hub and the mutual interference between fuselage 
and unit; this means that, in the aerodynamic design of a helicopter, allowance 
must be made for the parasite drag of the helicopter without rotor hub. 



260 



In this book, we use experimental data pertaining only to the aerodynamic 
design of the helicopter. Therefore, the graphs of the aerodynamic character- 
istics are given for conponents of. forces in a system of wind axes ty and t^ . 
The coefficients of forces t and h can tse obtained by the conversion fonnulas 
[eqs.(3.15) and (3.1?)]. 

The rotor angles of attack, plotted in Figs. 2. 142 - 2'll+5, correspond to 
zero deflection of the automatic pitch control mechanism. To reduce the voltmie 
of this book, the graphs of ty = f(a, Gq , V) or ty = f(mt, Gq , V), used for de- 
termining rotor pitch, are not given. 

METHDDS OF CONYERTING THE AERODYNMIC CHARACTERISTICS 

OF A ROTOR 

The conversion formulas presented below make it possible to use e^qjerimental 
data pertaining to some specific rotor for determining the aerodynamic charac- 
teristics of other rotors similar. to the tested rotor with respect to dimension- 
less geometric characteristics. For exanple, the experimental graphs of the 
aerodynamic characteristics shown in Figs. 2. 142 - 2.145 can be used, with the 
he3p of the conversion formulas, for determining the characteristics of rotors 
with other soUdity ratios if the rotors have rectangular blades, a twist of 
5-9°, and a profile close to the NACA 230 profile. The characteristics can be 
extrapolated to other rotors, but with a lower degree of accuracy. 

The use of the conversion formulas permits an appreciable reduction in the 
number of graphs of aerodynamic characteristics of rotors required for helicopter 
designs . 

3. Conversion of Aerodynamic Characteristics to a /245 

Different Rotor Solidity Ratio 

let us conpare two rotors whose blades have identical distributions of 

twist angles and relative chord b = -«-2 over the radius; the rotors are as- 

bo. 7 

sumed to have either a different nimiber of blades or a different chord' bo. 7, 
i.e., a different soUdity ratio. The magnitude of the mass characteristic of 
the blade y has only a minor influence on the rotor characteristics so that the 
difference in y can be disregarded; however, for rigorousness we will assume 
that Y of both rotors is identical. 

At uniform induced velocity distribution over the disk of these rotors, the 
flapping motion of the blades and all dimensionless coefficients in the body 
axis system- t, h, mt, and others - are identical if the rotors have equal 
values of the flight regime characteristic \j., collective pitch cp, and relative 
flow normal to the plane of rotation of the rotor \ : 



V sina — V 



--\j.tana — v:^^a — v. (6.9) 



261 



It is easy to prove this from the formixLas of the Glauert-Iock theory 
(Sect. 2) which indicate that the e35)ressions for all dimensionless coefficients 
contain foior quantities: |Jb, X, cp, and y These quantities fully determine the 
rotor kinematics; if they are equal, the velocity polygons in each rotor section 
will be alike, and the true angles of attack, Cy and c^p will be equal. 







Fig. 2. 147 Air Velocity Conponents Nonnal (V sin a - v) 
and Parallel (V cos a) to the Rotor Plane at Different 

Values of a and a. 
V cos ai ~ Y cos a^ ; V sin ofi - -^^ « V sin a^ - Vg . 



At known p,, X, and t, the angle of attack of the rotor is determined by the 
expression 



-+: 



ct 



tan<3.= h— =^ I /•■ 

(J. fi fi 4B2|Ji.-/|Jl2+ X2 



(6.10) 



Consequently, at equal dimensionless coefficients but different solidity 
ratios, the rotor angles of attack differ; the rotor with the larger solidity, 
i.e., with a larger dimensionless induced velocity, at equal t and at the same 
value of A., will have a larger more positive angle of attack. Figure 2.147 
illustrates the equality of the air velocity conponents normal (V sin o? - v = 
= \u)R) and parallel (V cos 01 = |iU)R) to the rotor plane at different mean induced 
velocities v and different rotor angles of attack Qf. 

Using the subscripts "1" and "2" to denote the quantities pertaining to /246 
rotors with solidity ratios of a^ and Cg > we can write the e^^jression for the 
difference of the angles of attack of both rotors 



tana, -tana^ = (°i-f2)<_^ 

4fi2^/(l2+)l2 



or, approximately for y, s 0.1$ at }? 4, |j,^. 



Act = C] — a2 = (oj — 02) 



4B2,i2 



(6.11) 



The difference in the angles of attack is expressed as the product of the 



262 



difference in the solidities and the ratio 



Consequently, at equal t 



and |J, the difference in the angles' of attack is proportional to the difference 
in CT. 



Thus, at equal |j, and cp all characteristics of the investigated rotors in 

body axes are identical if the rotors have angles 
of attack differing by a quantity Aa, which is 
determined from eq.(6.1l). To change over to 
characteristics in a system of wind axes, we use 
the formulas for converting from one system to 
another. Taking into accoiint that the difference 
in the angles of attack of the rotors Aor is 
small, we obtain 



Mrf -const 
V= const 
Mg- const 
ty 'Const 




Fig. 2. 148 Reconstiniction 
of the Dependence t^ = 
= f(mt) on Change of the 
Rotor Solidity Ratio. 



yj' 



'JTj 



iy^ Aa. 



Let us write out the final formulas for 
converting the aerodynamic characteristics of 
rotors with different solidity ratios: 



^1 = ^2-, 



yj 



^2' 



/ =i J_(o oj) ^; 



trif =m,. 



ai = a2 + (<'i-°2) 



4B2V/2 



(6.12) 



Equations (6.12) indi£ate that, on converting the characteristics, the 
value of the coefficients V, ty , and mt is retained whereas t, and a change by a 
quantity which is constant for given ty and V. This means that the reconstruc- 
tion of the aerodynamic characteristics of a rotor, represented as the depend- 
ence tjt = f(mt) at V = const and ty = const, reduces to a displacement of each 

curve along the ordinate by a quantity At, = Act ^ — (Fig. 2. 148)' 

4B^V^ 
Reconstruction of the characteristics need not be carried out in practice, 
since we can execute the aerodynamic design of a helicopter on the basis of the 
aerodynamic characteristics of a similar rotor with a different solidity ratio, 
with due regard for At^ . For exanple, in designing a helicopter with a solidity 
ratio CTg based on the aerodynamic characteristics of a rotor with a solidity 
ratio CTi , the required torque coefficient (i^t^.f ^s ^"^ ^^^ required power N^.f^ 
are determined in the following sequence: 



263 



a) We first determine f2Un 



t„ = 



^2 l/2 0{u./?)2/='o2 ' 

F,=— • 



b) For using the characteristics of a rotor vrith ai , we find 



c) From the characteristics of the rotor with a-y , we determine m^ , cfi , cpi . 

d) We find (m^ i)s, oi^.t^t ^.t„ s^d the required power of the helicopter: 

«A/2 = «i-(''i-°2)^; 

The formulas derived above were obtained for rotors with uniform induced 
velocity distribution over the disk. For rotors with an infinite number of 
blades and nonuniform induced velocity distribution, no argiunents or conversion 
formulas would change if at each point of the disk the induced velocities of the 
rotors with different soUdity ratios differed by an identical quantity equal to 
VAof. In reality, the change in solidity ratio influences also the variable com^ 
ponent of the induced velocity, i.e., the induced velocity diagrams do not differ 
by the same quantity. However, since we are converting the average per-revolu- 
tion characteristics ty, t^, and m^, which are mainly determined by the average 
portion of the induced velocity, the conversion formulas (6.12) can be used with 
stiff icient reliability. 

The proposed method of converting aerodynamic characteristics of rotors is 
similar to an analogous method for airplane propellers. It is also based on de- 
temiining regimes in which the kinematic characteristics of the propellers with 
different solidity ratios are identical. The difference is that, for propellers 
with different a, different flying speeds (different \~) are chosen, whereas 
for rotors for which the kinematics is determined not only by the normal veloci- 
ty but also by the velocity conponent in the plane of rotation, the flying speed 
is retained but different angles of attack are chosen. 

264 



4« Conversion o f Aerodynamic Characteristics on Variation 
in Minim um Profile Drag Coefficient of the Blade 
Sections c. 



If the blades differ in magnitude of the minimum profile drag coefficient 
[different quality of manufacture of the profile (see Sect .4.3), different /2i?i-8 
profiles differing mainly in c^pg], the following formulas can be used for con- 
verting the aerodynamic characteristics of the rotors. 

The increments in the coefficients of torque ajid longitudinal force, accord- 
ing to the Glauert-Lock theory, vdll be equal 

to 
W// = const 
V=const . Ac, 



M„= const 
ty= const 




Affi( = 



f-xp 



(1 + 1^^); 



AA=- 



Ac 



xp 



^, 



(6.13) 
(6.14) 



where Ac 



xp 



= (Cxp„)2 - (CxpJl • 



Po' 



Increased 



The remaining coefficients in the body 
axis system remain unchanged. The coefficient 
tx increases by an amount equal to about Ah. 
Thus, the conversion formulas will be 



V2-Vt; 



^2" 



Fig. 2. 149 Reconstruction of 
the Dependence t^ = f(mt) on 
Variations in the Minimum 
Profile Drag Coefficient of 
the Blade Sections. 



yr 



^-2 = ^x,+4-Ar,,V-; 



?2 = ?l. 



(6.15) 



The variation in the rotor characteristics •upon conversion is illustrated 
in Fig.2.1!t.9. 



ty t. 



Conversion can also be performed, provided there is constancy of the quanti- 

^2=^yi' 

ix =tr ■ 



Under this condition, the increment in torque will be equal to [see 
eq.(3.7l)] 

A m^ = Am;,r =^ Ac^p ( 1 + 31/2) 



265 



and, accordingly, 

'nt,=mt^-\-~Ac,^(l+3V'). (6-16) 

By analogy with eq.(3«72), in place of eq.(6.l6) we obtain 

Using eqs.(6.l6) or (6.I7), we change from point 1 in Fig. 2. 149 to point 3. 

The nonrigorousness of the conversion formula lies in the fact that, while 
it does take into account the effect of an increment in Cj^p , it disregards the 
fact that, at points 1 and 3, the angles of attack of the rotors are not mutual- 
ly equal (a^. = 0/2 ; cxs < ffg ) . Consequently, in these regimes there is a differ- 
ent distribution of the true angles of attack of the sections and therefore dif- 
ferent induced and profile powers at identical profile polars. However, the /2L9 
conversion methods are approximate so that the indicated inaccuracy is of no 
practical value. 

If the polars of the profiles differ not only in the quantity c^p but also 

in their slope, the proposed method will not be valid. Therefore, it is unsuit- 
able for converting aerodynamic characteristics to other Re and Mq numbers. In 
these cases, the quantity Am^ should be determined with consideration of the real 
values of a^ and c^p at each point of the rotor disk. 

5. Conversion of Aerodynamic Characteristics on Variation 
in the Peripheral Speed of the Rotor (Mq Numbers) 

Fig-ures 2.80 - 2.88 in Section 3 give graphs for the increment in torque 
coefficient at Mach numbers greater than 0.4* The graphs show that at moderate 
values of the thrxist coefficients the conpressibility of air has a noticeable 
effect on the quantity mt at Mq greater than 0.55 - 0.6. Therefore, vpon a 
change in Mq beyond these limits (and at near-separation values of ty at lower 
Mq) corrections must be introduced into the aerodynamic characteristics of the 
rotor. These corrections are determined from the graphs in ELgs.2.80 - 2-88 as 
the difference of the values of Amoo at the Mach numbers in question. 

For example, if the experiment was carried out at Mq equal to Mq^ , and the 
experimental data are used at Mq equal to Mq , then the value of m^ = mt(V, 
ty, tjt, Mq ) found from the e^qjerimental graphs, must be stpplemented by 

8m,„ = Am^^(Mo^)-Am^^(Moj), (6.18) 

where Amoo a^re determined at corresponding Mq numbers and at the same values of 

Thus, 
266 



mtiV, ty, tjc, ^02)='"^;,^ (^. (y, tx, Mo,) + Sm,:,,. (6 .19) 

At V > 0.3 when Mq changes, we must introduce a correction to the angle of 
attack of the rotor and, accordingly, to the blade pitch. These corrections are 
introduced analogously: 

a(^, ty, tx, ^^)^^<iyp (^. ^y. ^x, Mo,) + K^; (6.20) 

8a^<,=Aac^(^, ^y. ^jt, Mo2)-Aa iY Jy,t^,Ni^^. (6.2l) 

The blade pitch is determined in relation to V, ty, and 01 from the graphs 
of ty = f(ff, 9o , V) on the assung^tion that this dependence does not change with 
respect to Mq. 

6. Conversion of Angle of Attack and Rotor Pitch on 
Variation in Inclination of the Automatic Pitch 
Control. Flapping Compensator, and Mass 
Characteristic of the Blade 

As shown in Section 2, a change in the slope of the automatic pitch control 
and the flapping conpensator will not cause a change in the coefficients ty, t^, 
and m^ provided that the equivalent rotor angle of attack cv, = a - cpx remains as 
before. Consequently, conversion of the rotor characteristics reduces to find- 
ing the new rotor angle of attack by means of the expression 

a2 = ac+(?i)2 = ai- (?i)i + (9i)2. (6.22) 

where 91 is determined by the formula /250 



Cpi = — DjX -|- D{(\ — k 



l+ft2 



The coefficients ai and bj. are found from known values of ax and bx by 

means of eqs. (2-273), (2.274), (2.249), and (2.250). The rotor pitch is con- 
verted by the formula 

6o2=?a + Mo=S~"')(*'~*2)- (6.23) 

On variation in the quantity —^ — , it can be considered that the coeffi- 

cients ao and bj^ entering eqs. (6. 22) and (6.23) vary in direct proportion to 

v 
the ratio of the new and old values of -^ l -, x.e,, 

"•inn 



267 



In conparing eD^jerimerital data with each other or with calculated data, one 
miist also account for the effect of the rotor hub, as is done in the exaiiples 
given below. 

?• Examples of Using the Conversion Formulas 

Comparison of calculated aerodynamic characteristics with espjerimental . 
The calciilated aerodsmamic characteristics of a rotor with rectangular blades 
a = 0.0525 were obtained by the method presented in Section 4* The only differ- 
ence between calcxilation and esgjeriment, which should be taken into account when 
making a conparison, is the effect of the rotor hub on the esqaerimental charac- 
teristics. Taking "c^^ = 0.0015, a reduction of the calculation to the e^qjeri- 

mental conditions requires the addition of the following increment to the calcu- 
lated propiiLsive force of the rotor: 



or, in dimensionless form. 



^^-^=^-^^^ ^ 



A^,, =c; z!. (6.24) 



Thus, the e^^jerimental curves can be conpared with the quantity 

_ _|_ 2:0015 1/2^/ +0.02861/2. 
*ca/c ' 0.0525 'ca/c ' 

In Fig.2.150, e^qserimental curves and converted calculated curves {t^) are 
plotted for V = 0.3. The diagram indicates that, in powered flight regimes 

(tx = -0.01 0.02), the difference in m^ is negligible whereas, in autorota^- 

tion regime, the values of t^ differ by 5 - 15^ • The convergence of the experi- 
mental and calculated curves is better at lower V. 

Comparison of experimental aerodyn^ic cha racteristics. Using the experi- 
mental aerodynamic characteristics of a rotor with trapezoidal blades of /251 
NACA 230 profile, plywood planking, a = O.OS65, and Mq = 0.4, the conversion 
formulas will yield the characteristics of a rotor with rectangular metal blades 
with NACA 230 profile, a = 0.0525, and Mq = 0.5, which can then be conpared with 
the experimentally obtained characteristics shown in Figs. 2. 141 - 2.145 • 

The difference in the solidity ratio is taken into account by eq.(6.l2): 
<; (.=0.0525) = (' ('=0.0865) + (0.0525 - 0.0865) -^ = 

^tx (0=0.0865) — 0.009 =|- . 

268 




0.01 



-0.01 



-0.02 



-0.02 



> 


N 


h 


- 
























Experiment 

Conversion 




^ 




N 
y 


^ 






V 


















^ 


^ 


















- 


\ 


^\ 


N 


\ 


\ 
















\ 


v 


. 














w 


\ 


\ 


\ 














■ - 






\ 


w 


\ 


^ 




0.010 




mt 






\ 


s3 




V 












ty' 


?-^?d^ 




^ 












u.ia 
1^=0.10 


^ 


^ 




\ 




\ 








"=^ 




-^\^^ 

















1 










w 




\ 








., . 












- 






\ 




V 




V 


s 


\N 


k 




\\\ 


\ 


\ 














v 


V 


V 



K.g.2.150 Conparison of Experimental Iilg.2.151 Ccaiqsarison of Esqjerimental 
and Calculated Aerodynamic Character- and Converted Aerodynamic Characteris- 
istics of Rotor (Mfi = 0.15; V = 0.3; tics of Rotor (M^ = 0.15; V = 0.3; 



Mr 



0.5; cj = 0.0525). 



Mr 



0.5; a = 0.0525). 



The difference in the profile power of the rotors, with consideration of 
the difference in tlade planform, is found from the following expressions: 



m 



p>- 



,,=0.0525)=— ^(1+5K2)P,; 



nipr (0=0.0865) 



"xp. 



Ac, 



LJnpr = 



Ac 



xp 



■«*'(i+5p2)p+^^::^ 



'■xp. 



(l+5\/2)P; 



(6.25) 



(1+5K2)P ^ (1-1- 51/2) (p_p^) 



Admitting for the trapezoidal blade P = 0.92, c^p = 0.009 and in con- 
formity with the recommendations in Section 4«3, ACxp ~ 0.0025, we find 

Am^;.= -|'5^0.92+5i02? (0.92- 1)1(1 +5V^2)= -0.0004(1 +5K2). 

So as to make the conversion of the characteristics only with respect to /252 
one of the coefficients, namely with respect to t^ at mt = const, we convert 

269 



Anipi. to Atjt . 

M,^^^^. (6.26) 

The difference in the Mq mjmbers is taken into accoxont by eqs.(6.19), while 
Anioo is determined from the graphs in Figs. 2. 80 - 2.88. 

At a difference in solidity ratios, the hub drag resxilts in a different in- 
crement of the coefficient t^ : 



(^^J.=^'^^= 



Cj- 



To convert the characteristics, we must subtract (At, )x from the coeffi- 

hub 

cient of the propulsive force of the e^^jeriment under conversion (oi ) and add 
(^■tx....)2: 



*hut ■ 



'"■.>=''>-~^-{^i,r+{¥lr- ^'-^^ 



Since, in analyzing the experiment with trapezoidal blades, the drag of the 
nonrotating hub was excluded, we can take c^ = 0.00075 to account for hub /253 

rotation. Therefore, the conversion for the effect of the hub is performed by 
means of the expression 

. . / 0.00075 0.0015\t7, 

^ (.=0.0525) = /. (.=0.0865) -^-^^^^^^ ^^^^)V = 

= /:r (.=0.0865) +0.02\72. 

Thus, the final expression for converting the coefficient t^ has the form 

ix (0=0.0525) ^tx (0=0.0865) — 0.009 ^| 

_ 0.0004 -^^tj^ I 0.02 K2 4- ^ . 

For conparison purposes. Pig. 2. 151 gives the e^qjerimentally obtained and 
converted (t^) characteristics. For the most part, the agreement of the curves 
is satisfactory. 

Section ?. Performance and Propulsive Efficiency Coefficient 
of a Rotor 

The helicopter rotor produces lift and simultaneously acts as the prime 
270 



mover of the helicopter. Therefore, it is natioral to characterize its lifting 
and propiilsive properties in the same manner as a wing is characterized by the 
performance K„ and a tractor propeller by the efficiency Tlt.p • These concepts 
permit definition of the degree of siiitability of a rotor as a means for produc- 
ing lift and propulsive force, as well as a rapid performance, in general form, 
of approximate calculations of the required power of a single-rotor helicopter, 
a helicopter with a wing and tractor propellers, or a multirotor helicopter, and 
proper selection of the regime of maximum performance (maxim\m range). Knowing 
the performance and efficiency, one can estimate directly the expediency of in- 
stalling a wing and tractor propellers on a helicopter, determine what part of 
the total drag of a helicopter is made vp by parasite drag and how much the re- 
quired power can be reduced when the parasite drag is reduced, and find the 
rational distribution of power between tractor propellers and rotor. 

A determination of rotor performance in an autorotation regime is carried 
out in the same manner as for a wing. The concept of rotor performance has been 
widely used in aerodynamic designs of autogiros. The rotor performance, together 
with a coefficient which we will call the propulsive efficiency coefficient, can 
be used also for calculating a helicopter, as we will demonstrate below. 

Unlike in an airplane, where the wing and propeller are different units and 
K„ and \,j, can be examined independently of each other, in a helicopter the 
rotor performance K ajid the efficiency T] are interrelated and the efficiency of 
a iTotor in any regime is determined by the value of the product KT]. 

Let us first discuss the concepts of performance and efficiency, described 
in individual works on helicopter aerodynamics (K.Khokhenemzer and other 
authors) . 

1. Performanc e and Efficiency of Rotor Proposed /254 

by K.Khokhenemzer 

Rotor performance can be determined on the assimption that the actual pro- 

piilsive force of the rotor (-X) is the difference between the ratio -l and 

some arbitrary drag of the rotor Xaj.Tj\^-X = -^ X^rbj* from which we determine 

the arbitrary drag of the rotor: 

Correspondingly, the rotor performance is 

y r 

+x 



t/'^^^ 

Xars ~'75N . ^ ■ (7.1) 



V 

The ratio —^ — would be equal to the propulsive force if the entire power 
were converted without losses into propulsive force. Since the actual propulsive 

271 



force is equal to —^ ^art* ^'^ ^s oTdvIous that all losses belong to X^^ . 

Thus, the rotor is represented as a certain mechanism creating forces Y and X^ri, ; 
the power si^jplied to it creates, without losses, a propulsive force equal to 

■ 1^ so that the total (actual) propulsive force is equal to '^ ■ ■ - X^^^,, . 

The arbitrary drag of the rotor is conparable with the drag of a wing plus 
the power losses of the tractor prcipeller, i.e., 

while the total propulsive force of the system "wing + tractor propeller" is 

equal to ■ '^^y^'" -X„^^. 

As a second version it is proposed to consider that the lift of a rotor is 
produced without loss (without drag) and that all losses are accounted for by 
the generation of a propiolsive force. The rotor is represented as some mechanism 
producing lift Y, while the power sipplied to it is converted into propulsive 
force 



Hence, 






,,==^=1=^. (7.2) 

75iV m^ 



It is obvious that the efficiency T] cannot be conpared with Tlt.p but with /255 

the difference Tl^.p =— ^ -» since the thrust of the propeller minus the 

wing drag is equal to " 






'tE. 



The description of both versions of representing the characteristics of a 
rotor shows that they are both artificial and conparable only with combined 
characteristics of the wing and tractor propeller. This is the adverse side of 
the proposed concepts. Their favorable side is that the characteristics are de- 
scribed only by one quantity: either by performance or by efficiency. 

The concepts of rotor performance (or efficiency) examined above are con- 
venient for calculation, since they relate flying speed V and helicopter weight 
G (or propulsive force) with the required power. Actually, having set in 
eq.(7.l) G = Y and Qpar = -X, we obtain 

272 



^^•/=^(f+^.'> 



(7.3) 



However, the sense of applyiiig these concepts is predicated rpon the con?- 
venience of use in calculation and in determining optimal parameters. Otjt con- 
cepts presented below also simpHiy the calculations and, furthermore, while re- 
taining the sense and value of analogous con- 
cepts for airplanes, facilitate an investigation 
V- const Mo^const °^ conposite rotary-wing aircraft. 

e= const 
Y= const 

2. DeteiTiiination of Performance and Propulsive 
Efficiency of a Rotor 

Let the rotor operate in the regime "a" 
(Fig. 2. 152) with a propulsive force Xa, requir- 
ing a power N^. To increase the propulsive 
force to X^ while retaining the lift Y, the 
rotor must be tilted forward and the power must 
be raised to N^. The efficiency of the rotor as 
a propulsion unit on changing from the regime 
"a" to the regime "b" is defined as the ratio of 
the power increment of the rotor as a prime 
mover -AXV to the increment of power stpplied to 
the rotor: 




Fig. 2. 152 For Determining 
the Concepts of Performance 
and Efficiency of a Rotor. 



Tl = - 



— IlXV 



-At^V (t,^~t,^)V 



75\N 



Am». 



nif —ntf. 



(7.4) 



If, to increase the propulsive force of the rotor we were to install a 
tractor propeller and supply it with a power equal in magnitude to the differ- 
ence N^ - Na = AN, then it would create a thrust of 






(7.5) 



A conparison of this e^q^ression with eq.(7.4) shows that the incitement in 
propulsive force of the rotor -AX is characterized by its propulsive efficiency 
T] just as the thrust of a tractor propeller is characterized by its effi- /256 
ciency Tlt.p • 

In a craft with a rotor installed to produce lift, the power can be sxpplied 
either to the rotor (helicopter: Nt.p = 0, N^ot = %* see Fig. 2. 152), or to the 
tractor propeller (autogiro: N^o^ =0, N^.p = N,,), or distributed between the 
first (Ni.ot = Na) and second (N^.p = N^ - Na). A comparison of T\ with Tlt.p 
shows which of these versions is better, i.e., whether it is expedient to install 
a tractor propeller for increasing the propulsive force of the craft or whether 
it is more advantageoiis to transmit the entire power to the rotor: if 1) > Tlt.p, 



then |Xi,| > |Xa| + Pt.p » More precisely, we must ccaipare 11 with Tlt.p 



§ 



VjjE. 



273 



(the ratio *'P takes into account the difference in losses of power trans- 
mitted to the rotor and to the tractor propeller) • 

Thus, in order to obtain for a rotor, which is a lift-producing coriponent 

as well as a propulsion unit, a coeffi- 
cient analogous to the efficiency of a 
y tractor propeller, it is necessary to 

investigate the increment in propulsive 
^ _v. force (or drag) of the rotor when power 
^ — is supplied to it. Therefore, we de- 
^ ' 1 fined the propulsive efficiency of a 



-X 



N 
■ Direction of flight ^ rx-x)y Totor 33 the ratio of the incromonts of 

• " /(=-i- Tj= ^ <^ '^ useful and expended work, although such 

'"''^ ^^'^ a ratio is not actually the efficiency 

but only performs its role for craft 
Fig. 2. 153 Model Representing a with a rotor. 
Rotor as Two Elements - lifting 

and Propelling. To determine the propulsive effi- 

ciency with respect to eq.(7«4) we must 
select some regime as the initial (we 
select here the point "a"). The drag of the rotor in this regime X^ determines 
its performance. 

It is ej^edient to take, as initial regime, an autorotation regime (point 
"c" in Fig. 2. 152). in this regime, no power is supplied to the rotor which, in 
producing lift, also creates drag like a wing. 

Thus, the work done Ipy a rotor can be interpreted as follows: The lift is 
generated by the rotor in an autorotation regime without the expenditure of 
engine power, just as for a wing; in regimes with a supply of power the rotor 
creates a propulsive force which partially conpensates (at N < Nj,) or overcomr- 
pensates (at W > Nj^) the rotor drag in an autorotation regime. The propulsive 
efficiency characterizes the power losses of a rotor when changing to an engine 
(prcjpulsion tonit) regime. The rotor is replaced by the model shown in Fig. 2. 153, 
for which, in conformity with the foregoing, the expressions for T] and K have 
the form 

■^- ,n, ' (7-6) 

••c 

In eqs.(7.6), (7»7) and below the subscript "c" means that the indicated 
quantity refers to an autorotation regime. 

In level flight, the propulsive force of the rotor is equal to the sum of 
the rotor drag t^ and the parasite drag of the helicopter t^ - t^ = t^ + /257 

+ "Cjj ~— — . Consequently, the propulsive efficiency and performance of a heli- 
copter in horizontal flight are equal to 
274 






K, 






or 



Ku 



GV 



h,-^ 75Ar,_^r, 



(7.8) 
(7.9) 

(7.10) 



We note that the quantity (t^ - t^ ) represents the coefficient of the 

arbitrary propulsive force of the rotor in horizontal flight and is equal to the 
coefficient of the drag counteracted by the tractor propeller of an autogiro or 
helicopter for which, in horizontal flight, the rotor operates in an autorota- 
tion regime. 

V = const, M„-const 
e = const 
ty= const 

U=const Mg=const 
6= const 
tyconst 




%^ i^="«^ 




K 



V*'. 



V 



Fig. 2. 154 Determination of Rotor 

Efficiency at linear Dependence 

of tx on m^ , 



Fig. 2.155 For Estimating the Expedi- 
ency of Installing a Tractor Pro- 
peller on a Helicopter. 



In eqs.(7.7), (7.9), and (y.lO), the quantities K and K^ are the same as 
those used in autogiro calculations. 

Use of the concepts of performance and propulsive efficiency for calcular- 
tion is highly convenient in the case of linear dependence of t^ on m^ . Here, 

the quantity 71 does not depend on m^ or t^, since the ratio — jr- is equal to the 

angle of slope of the straight lines t^ = f(mt) (Fig. 2. 154). 

In place of the aggregate of the graphs (see Figs. 2. 105 - 2.109) constructed 
for several V, the aerodynamic characteristics can be represented as two graphs: 
K and 7] as a function of ty and V (see Fig. 2. 158), by means of which t^ is 



275 



determined from eqs.(7.7). and (7.6) at known values of ty, V, mt, or else m^ is 
determined at known ty, V, t^^. 

In the case of a nonlinear dependence of t^^ on mt, when T) depends on mt, 
the use of K and T) in calculations offers no substantial advantages. Here, it 
is of interest to determine the propulsive efficiency with respect to the /256 
angle of inclination of the tangent to the curve t^ = f(mt) at the point in 
question 



^1 = 



dm J, 



(7.11) 



A conparison of Tli with Tlt.p i'^ permits determining whether the propul- 
sive force of a craft can be increased by installing a tractor propeller. Let 



V'ConstiMg'Const 
i^=const 




■f]^ = 7]^^^ 5^.T> at the point "a" 

It is obvious that, if 



(Fig. 2.155). 

mt < mt^, we have Tli > Tit., 



%LL- and use 



^=-' 



Fig. 2. 156 For Determining the 
Performance and Efficiency, at 
Nonlinear Dependence of t^ on mt > 



of a tractor propeller is not advantage- 
ous. If the rotor operates in a regime 
with mt greater than mt , the installa- 
tion of a tractor propeller may increase 
the propulsive force, the maximum gain 
being obtained when a power corresponding 
to mt is transmitted to the rotor and 
to the tractor propeller (mt - mt ) . 



In the case of a nonlinear dependence 
of tj on mt , it is preferable, for ap- 
proximate calculations with the concept 
of performance and efficiency, to replace 
the nonlinear dependence of t^ on mt by a linear dependence. Such an approxima- 
tion is made in the segment of the straight lines from a = -20° and sometimes 
from a = -15° at V = 0.15 (point F in Fig. 2.156) to the minimum value of mt 
(point H in Fig. 2.156) at which the greatest deviation of mt from exact values 
does not exceed 3% of mt at t^ =-0.1 V^ (approximately a horizontal flight 

regime of helicopters) . 

The value of T] and K, determined from the approximating segment HF, is 
calculated by the formulas 






V 



K=^=- 



mt^ 



(7.12) 
(7.13) 



276 



The efficiencv detennined by the angle of Inclination of the linearized de- 
pendence tjt = f(mt; must be regarded as the propulsive efficiency, on the aver- 
age, for the cturve. 

3« Performance and^Efficiency of a Rotor, Obtained 
from Experimental Data 

The graphs for the aerodynamic characteristics of a rotor in the form of 
the dependence t^ = f(mt) obtained from experiments in a full-scale vdnd tunnel 
are given in Section 6 in Figs. 2. 142 - 2.145 • They pertain to a three-blade 
metal rotor with rectangular tvdsted blades, a = 0.0525, and include the drag of 
the rotor hub. From these graphs and using eqs.(7.6), C7.7), we determine the 
dependence of K and 11 on ty shown in Figs. 2. 157 and 2.158. 




n 












w 








= 


- 




















■ 






— 


■^ 








-^ 


. 












i 




t 






iZ^^J 


no 












. ^=0.75 
n?.? 


^ 


































0.84 
















t 


>F- 


■R?0 



0.20 ty 



0.10 



ais 



OJZO ty 



Fig. 2. 157 Rotor Performance according Fig. 2. 158 Rotor Efficiency according 

to Experimental Data (Mq = 0.5; to Experimental Data (Mq = 0.5; 

a = 0.0525). o = 0.0525). 



To define the character of the slope of the curves of K, let us examine /259 
the approximate expressions for K. According to the energy method of calcula- 
tion (Sect .3), we have 



1 






V 

1 



f<ind 



\pj. 



'^ind 



^Pr. 



K 



V 4V2 



(7.14) 
(7.15) 

(7.16) 



pr 



277 



The increase in performance with an increase in V can be attributed to a 
decrease in induced and profile drags with an increase in V. At average values 
of ty, the performance depends little on ty since the induced part of the re- 
ciprocal performance increases with increasing ty whereas the profile part de- 
creases (tp to incipient flow separation). At small ty, the performance de- 
creases owing to an increase in 



Kpr 



The maximum magnitude of rotor performance depends on V, Mq, ct, quality of 
blade maniifacture, and geometric blade characteristics . Optimum performance was 
not obtained in the experiments, owing to the low value of Vj,ax • The largest 
of the obtained values is K„ax = 9*7 at V = 0.3, ty = 0.1?- 

Rotor performance is lower in magnitude than wing performance. This is ex- 
plained by the fact that a rotor has greater profile losses than an airplane 
wing since, at equal flying speed, the flow across the blades has a much greater 
velocity U. In the case of undeflected mechanisms, the profile drag of a wing 
is by a factor of 2 - 2.7 less than that of a rotor. Upon deflection of the 
mechanisms, the profile drag of a wing increases appreciably and approaches the 
profile drag of a rotor. 

A rotor and wing are closely adjacent in value of induced drag (at t„ = D 
and at uniform induced velocity distribution, the induced drag is the same) . 

A decrease in performance at small V is inevitable both for a rotor and /260 
for a wing, owing to an increase in induced drag. However, the wing cannot have 
as low a £^erformance as a rotor, since the wing cannot have as high a Cy as a 
rotor at V < 0.15 [see eq.(4.37), Chapt.IIl]. 

The propulsive efficiency of a rotor varies within the limits of 0.99 to 
0.9* The curves of T] intersect one another, and in some cases there is an ap- 
preciable scattering of the test points. The fact is that it is difficult to 
determine accijrately the quantity T], since the scattering of the test points on 
graphs of t^ = f(mt) creates some indeteradnacy in the angle of slope of the 
straight lines, which has a noticeable (within 3 - 5%) effect on the quantity T\. 

The inaccuracy in determining T], and also K, shows that when representing 
the characteristics of a rotor in the form of lifting and propelling elements it 
is inpossible to estimate them separately with high accuracy. However, this 
does not mean that calculations performed with the use of K and T] have a low ac- 
curacy, since when determining m^ by the formula 



(^-'-) 



V 



„ _, .__ (7.17) 

the errors in determining T] and K are conpensated . 

Figure 2.158 shows that, even if the low accuracy of determining the rotor 
efficiency is taken into consideration, this efficiency is greater than that of 
a tractor propeller. Since 7] is defined as the ratio of the increments of use- 
ful to e353ended work, it need not be less than 1.0. We will e^gjlain this. let 

278 



us substitute into eq.(7»6) the expression for rotor power taken from the 
energy method of calculation 

75N=15Npr-^15Ni„j~XV. (7.18) 

Then eq.(7«6) takes the form 



T] = 



(Xc -X)V + 7h {Nm - W,We) + 75 (V-^^) " ( 7 -19 ) 



It is clear from eq.(7»19) that, if the induced and profile powers of the 
rotor which depend mainly on lift were not to change on a variation in the pro- 
pulsive force, then the propulsive efficiency wotild be equal to 1.0. Actually, 
the differences of N^^ and Np^ are small, since we are examining the change of 
propulsive force at constant lift and flying speed, i.e., at approximately 
identical average values of induced velocity and true angles of attack of the 
blade sections. We can show that, for an identical propulsive force, these dif- 
ferences are respectively smaller than the induced and profile powers of a 
tractor propeller, as a result of which T] > llt.p • 

Thus, in examining the lifting and propulsive properties of a rotor, we de- 
termined that the bulk of power losses (Nj^^ and Np^ in an autorotation regime) is 
accounted for by energy losses related with the production of lift, which deter- 
mines the low performance of a rotor. 

The propulsive efficiency of a rotor differs from 1.0, owing to the small 
difference in induced and profile losses in regimes with power supply to the 
rotor and in autorotation regimes; it is greater than the efficiency of a 
tractor propeller. 

It should be borne in mind that the values of K and Tj, whose dependence /261 
on ty is shown in Figs. 2. 157 and 2.15S, are valid for regimes within limits in 
which the e^qDeriments are carried out. This means that mt, calculated by 
eq.(7.17), can be correctly determined, if it is not greater than the maximum 
values of mt up to which the experimental curves were plotted (mt 
= 0.01 - 0.013). "*'' 

4. Performance and Efficiency of a Rotor. Obtained 
from Calculated Graphs 

The performance and efficiency of a rotor with rectangular twisted blades 
(variant II in Table 2.10), a = 0.091, were determined from graphs of the aero- 
dynamic characteristics obtained by calculation. In the case of nonlinear de- 
pendence tj = f(mt), the quantities K and T) were found from eqs.(7.l2) and 
(7.13). The graphs of K and 1] are shown in Figs. 2. 159 and 2.160. 

Rotor performance begins to decrease at Mg > 0.6, especially at large Vj at 
Mo = 0.7 and V = 0.3, K diminishes by 1.5, and_at V = 0.4 by 3.5. At Mq = 0.7, 
the performance at V = 0.3 is greater than at V = 0.4, and the maximum performr- 
ance is equal to about 7.5. 

279 



The efficiency of a rotor, for Mq =0.6 - 0.7 at average and small values 
of ty, has a higher value (more than 0.95). At near-separation values of ty, 
the efficiency begins to drop markedly, but does not decrease when ty = ty 

less than 0.75 - 0.85. 



K 


■^ 






















^ 




_ 




































M 


-.nil 




" 


















^^ 






v=u.w 
























W 
















,- 






^ 


^ 




















1 


%=ft6 


9 












, 




^ 




























^ 


_ 




J 




M,=0.7 








_, 






^ 
























~ 


— 


■- 


8 






/ 




^ 


' 








^ 






























Y 
V'0.30 
















/ 


^ 








^ 


































^ 














7 


/ 


'/ 




^ 


















^ 


1 — ' 


i— 


I—' 






" 


"" 


-- 


-^ 


/ 
















C^ 
















^ 


- 


"" 




V'OAO 






































6 












-> 


'' 










^ 










































,^ 


^ 










v 


^ 




-J 


J 




















M 'QAandO.G 


5 


^■ 


r 








,_j 


^ 






[ 




















, ,., 


~~ 






1 1 


r— n 























U- 


^ 


"^ 


-- 


- 


" 
































'- 


-■ 


— 


~, 


.^ 




^ 






' 1 


f 


i> 


' 




























- 


























/=fl20| 





t 


^ 


^ 






« 






^ 






^, 




,M= 






3 


- 




— 


- 




— 


- 






— 


- 


— 


— 






— 


- 




- 


- 


- 










"=^^ 


;;— 4- 


2 








































IF n.„rr\ 1 
















































_ 


_ 


¥ • 


-U.IO' J 



0.08 0.10 m 0.1 1 0.16 0.18 0,20 0.22 0.21 ty 
Fig. 2. 159 Rotor Performance (Calculation, a = 0.091). 



The values of K and Tl obtained as a resiolt of linearization of the curves 
of t^ = f(mt) hold true within certain limits. The tpper limit of applicability 

mt / 7CK 
of the graphs of K and T] are the values of the ratio — "ii'- = V~YV 
given in Table 2.13- "^ 

7 
1.0 



a95 



0.90 



ass 



0.B0 



fl7J 



~ 








z 


11 


— 






"^ 


^ 








■r- 




— 


^ 


-- 






^-. 




— 




.^ 


I' 


■_ 




r= 






— 


-~ 


~^ 
















L. 


L 


-t- 


— 




■ — 


-- 






>=o 




— 




-• 








■ 


— 






_. 


w^o 


.7 


















zo' 


- 




p. 


-=: 






~ 




' — 


— 




L. 


Ij;; 


■- 


-^ 






Vmni, 








■^ 














^ 


Sr 


ti 




■^ 


~" 


*^ 


E^ 




— 




— 


■^ 


k-i 


*^''- 














<5 




■v 

^ 

K 


i _ 


\ 


\ 


\ 


\ 










^ 


^, 


^' 


^ 


M 






r-- 


«5 




».^ 








<^ 












* 


h> 






>.■ 






\ 




"^ 


^ 










-< 


s 


















\ 




Xx 


■V 






s 


s 






^ 




















> 




\ 


\ 


\ 






\ 








s 




\ 


^ 






















\ 


" 




\ 






s 


^ 


























\ 






> 






\ 




























c 


\V 






N 






s 






\ 






\ 


N 
























\ 


% 






s 






N 


i 






\ 




\ 






























- 


- 


\ 


V 


_ 


\ 


^ 




\ 


\ 




\ 


\ 


\ 

1 




^^-^ 






























































\ 


^\ 


\ 


t 


































- 




\ 














\ 




\ 












































\ 


- 


.^\ 




^ 










































- 











\ 




.s> 






























VXr 


I 


I 








\ 


^\ 






^ 








M„=0.6 
























* 


\'- 




























































M =07 
























\ 


































• X 






































_ 


^J 










OM 0.10 012 p.l't 0,16 0.18 aZO 0.22 O^/fty 
Fig. 2. 160 Rotor Efficiency (Calculation, o = 0.091). 



280 



If the ratio 



75N 
YV 



obtained from calculation is less that that given in 



Table 2.13, the calciolation will not differ by more than 3% from the calctalation 
made from graphs of aerodjmamic characteristics. Furthermore we note that the 

75N 



values of 



YV 



given in Table 2-13 correspond to flight regimes with a = -20° /262 



(sometimes -15° at V = 0.15). 

The lower limit of applicability- of the graphs in Figs. 2. 159 and 2.160 is 
the autorotation regime or a powered glide. More accurately, the calculations 
for an autorotation regime are performed from the graphs of K^ shown in 
Fig. 2. Ill, since the values of K found from the linearized curves may differ 
somewhat from K^ . 



TABIE 2.13 



/ 75/V \ 
V YV ) 



0.15 



0.6 



0.2 



0.48 



0.3 



0.37 



0.4 



0.3 



A Gonparison of the performance and efficiency of rotors having blades with 
different geometric characteristics shows that, at Mq = 0.4 - 0.5, the blade 
profile influences the value of K to within several percents whereas T] depends 
little on the blade profile. This means that rotors with different profiles re- 
quire a power differing by AN, where AN is independent of the type of operating 
regime of the rotor, namely at either large or small propulsive force (in gliding 
or climbing). 

For trapezoidal blades (variant I of the blades in Table 2.10)_j_ K is greater 
by 0.5 (at V = 0.2) - 1.5 (at V = 0.4), and T] is lower by 0.01 (at V = 0.2) - 
- 0.03 (at V = 0.4) than for rectangular blades. This means that the greatest 
decrease in required power for a rotor with trapezoidal blades occurs at small 
propulsive forces. At large propulsive forces, the rotor with rectangular /263 
blades having a larger 1\ may prove to be better. 

Conparative graphs of K and 11 at Mq = 0.7 are shown in Figs. 2.161 - 2.164. 
The diagrams show that, for a rotor without a high-speed profile at_the blade 
tip (variants III, IV), K is smaller by 0.7 (at V = 0.2) - 1.7 (at V = 0.4) than 
for a rotor with a high-speed profile. For trapezoidal blades, K is higher by 
0.5 - 1.0 and T| lower by 0.02 - O.OS, respectively, than for a rotor with rec- 
tangiilar blades. 

For a rotor with blades of increased geometric twist (variant VI) and with 
blades e^qjanding toward the tip (variant VIl) at V = 0.4, the performance is 
0.5 - 0.7 lower and the efficiency 0.05 - 0.15 higher. The very high value of 11 

281 



J\ 



for a rotor wLth increased twist is a consequence of the substantial decrease 
in profile losses tpon an increase in propulsive force of the rotor (see 
Pigs. 2. 75 and 2.76). 





- 










_ 





1 


_ 












7 ^ 




- 


























^ 




— 






















































■^ 


ill 
1 


pj 






























































0.W 0.12 0.11* 0.16 



0.18 



0.10 0.12 O.l't 0.16 0.18 ty 



Fig. 2. 161 Performance of Rotors with Fig. 2. 162 Efficiency of Rotors with 
Blades of Different Shape (V = 0.2; Blades of Different Shape (V = 0.2; 
Mo = 0.7). Mo = 0.7). 



K 




' — 


— 












^- 


-- 


1 


1 























/? 


6 






^ 












^ 


-- 


m 




' 








^ 








-- 






1 


-■ 






•-- 




-- 






5 




-^ 


---' 


^ 




" 




--- 


- 


— 














" 







111 






■—■ 


' 






' 












■s 

























1 


~~ 










^ 


X 


^ 




V 






^ 


\, 




- 








































: 


^ 


^ 












s 


1.0 


-— 






■X 










"- 




_. 


r: 




- 


- 


- 


030 


- 


— 


' 






























\ 


nsn 
















^ 



VII 



aw 



ai2 



W* ty 



aw 



0.12 



Fig. 2. 163 Performance of Rotors with Fig. 2. 164 Efficiency of Rotors with 
Blades of Different Shape (V = 0.4; Blades of Different Shape (V = 0.4; 

Mo = 0.7). Mo = 0.7). 

5. Conversion p_f Pe rfggngi-Iice and Efficiency on 
Variations in^Rot or Parameters 

In conformity with the formulas deri-ved in Section 6, the performance of a 
rotor on variations in the soUdity ratio and profile power coefficient is con- 
verted by the expression 



1 (a — oj) ty Am^yti 



(7.20) 



/< /Ci 



4B2K2 ' tyV 



282 



A change in profile power coefficient should take into account a change /264 
in Cxpq of the profile (or Ac^p , owing to the difference in the quality of blade 
manTrCacture) and a change in mp^ from the wave drag: 



Mn. 



Ac_ 



•(l+5l/2)P + 5/n„ 



(7.21) 



The propulsive efficiency is independent of the difference in a and Ac^p . 
The quantity bva.^^ depends on t^ so that also T] depends on t^ •• However, for the 
sake of singjUcity w6 need not convert T], and we substitute 5moo into eq.(7.2l) 
at an average value of t^ • 



6. General Conim ents on Rotor Efficiency and Performance 

Figures 2.165 - 2.16? show the generalized graphs of K and T], which can be 
used for estimate calculations. Figure 2.165 gives the graph of K for a = 
= 0.091, which is valid for average and_large ty . At small (ty ~ O.l), K is 
smaller by 0.2 (at V = O.I5) - 1.5 (at V = 0.4). Figliore 2.166 contains the 
graph of 11 used for all ty at Mq ^ 0.55- For Mq > 0.55, the efficiency must be 
corrected by a quantity AT], which is plotted in Fig. 2. 16? as a function of ty. 



M, 



o> 



V. 




0.1 



Rectangular blade 

Trape zoidal blade 

LI I \'\ 1.1 rm I 

0.2 0.3 




0.8 



~ Trape zoidal blade 

li Jill rn jjn 



0.1 



0.2 



a3 



Fig. 2. 165 Generalized Graph of Rotor Fig. 2. 166 Generalized Graph of Rotor 
Performance (a = O.O9I). Efficiency (Mq s 0.5). 

Thus, as shown in Figs. 2.165 - 2.167, the rotor performance is lower than 
the wing performance, and the propulsive efficiency of the rotor is higher than 
that of a tractor propeller. This is e^iplained by the fact that the bulk of the 
power losses pertain to losses related with the production of Hft, whereas the 
propulsive efficiency differs from unity owing to the small difference in in- 
duced and profile losses in regimes with power sipply to the rotor and in auto- 
rotation regimes. 

Thus, it is obvious that the installation of a wing with a performance 
higher than that of a rotor will increase the performance of the lifting system 
of a helicopter. The installation of a tractor propeller of an efficiency lower 
than the propulsive efficiency of a rotor will lead to some increase in required 



283 



/>'?|0.T0 0.1Z O.l'f 0.16_ 0J8 0.20 0.22 ty 




power. Therefore, a tractor propeller on 
a helicopter can be usefiol when the relief 
of the rotor load by the wing or the /265 

reserve of available power render the 



ratio 



75N 
YV 



greater than that shown in 



Fig. 2. 167 Correction for Rotor 
Efficiency as_a Function of ty. 
Mo, V. 



realization of a large power excess 
> 0.45 - 0.5 for a helicopter. 



Table 2.13, since then the negative angle 
of attack of the rotor becomes greater 
than 20° (which is undesirable for design 
considerations, since the range of pitch 
angles of the helicopter and its parasite 
drag will increase) . Furthermore, at 



larger 



75N 
YV 



the values of T) may become 



smaller than Tl^.p . A tractor propeller or 
another propeller may be required for 



approximately -~- > 0.35 
G 



when V„ 



> 



Quantitatively, the change in required power of a helicopter, on installa- 
tion of a wing or tractor propeller, is small. Such an estimate will be made in 
Section 4, Chapter III. 

Section 8. Calculation of Rotor Characteristics inHovering 

and Vertical Ascent^ (Mom entu m Theory of Propellers ) 

The theoiy of a rotor in hovering and vertical ascent has been thoroughly 
presented in the literat-ure on helicopter and propeller aerodynamics. In this 
Section, we will give some data pertaining to a calculation of rotors with peri- 
pheral speeds as they are in use at present. 

The calciilations were performed with regard to momentum theory of a rotor. 
This theoiy was selected because of the fact that introduction of linearized 
aerodynamic characteristics of the profile into the calculation can be replaced 
by introduction of the actual dependence of Cy and c^p on a and M, obtained from 
wind-tunnel tests of the profile. 

1. Brief Review of the M_o mentum Theoir of Propellers 

Figure 2.168 shows the velocity polygon in a blade section at a relative 
radius r in the regime of vertical climb. The resultant velocity of flow in the 
blade section U represents the sum of the vectors: flying speed Vy , peripheral 
velocity lur, and induced velocity u. Since the vector of the resultant aerody- 
namic force of the section dR is directed opposite to the momentimi vector, the 
induced velocity vector u is parallel to dR. 



284 



J 




Fig. 2. 168 Velocity of Polygon in a 
Blade Section in Vertical Climb-Regime. 



The mass flow through a /266 

circular section at a radius r of 
width dr is equal to 



dm = Q2nrdr\Vi\, 



(8.1) 



where Vi is the vertical velocity 
conponent U. 

Applying the momentum theorem 
to the ring and using the theorem 
of doubling the induced velocity 
far aft of the rotor, we obtain the 
equation 



2dmu=Zf^dR, 



(8.2) 



where z^dR is the resultant of the elementary aerodynamic forces created by all 
blades at radius r. 

Substituting eq.(8.l) into eq.(8.2) and expressing dR in terms of the force 
coefficient Cr -^q obtain ^^^ equation 



4jtq 1 1/, I «r afr = z^^c^^^^^q — b dr. 
Equation (8.3) can be represented in the form 



(8.3) 



|1/Jk = 4-6c„ 6/2. 
" 8r sec. 



(8.4) 



This equation determines the relation between the velocity of the air and 
the coefficient of aerodynamic force in the blade section (the so-called 
"coupling equation"). 

Equation (8.4) holds for a stream flow throvigh the rotor disk but is in^ 
applicable in the region of the "vortex ring" . 

Expressing the velocities entering eq.(8.4) in terms of trigonometric func- 
tions of the angles of the velocity polygon, we can write eq.(8.4) in the tri- 
gonometric form: 



sinp-sin(Po — P) of 

•=— ^ Co 

COS (po — P + v-fr) 8r '^sac. 
where p is the inflow angle. 

The quantities |j,pr and Pq are equal to 



(8.5) 



(8.6) 
285 



%=tarr'-^-. (8.7) 

On the left-hand and right-hand sides of the equations, Pq is a known 
quantity at a given r; p, ofj., and ^p^ are tmknown. However, if we assign of, , 
then the characteristics of the profile will yield lip, and Cr^ , while 
eq.(8.8) will furnish g: 

p = a, — 9 = a,-(eo-*ao + A?)- (8.8) 

The problem consists in determining a, at which eq.(8.5) is satisfied. This 
value of ffj. is found by successive approximations. Simultaneously with deter- 
mining Qfj. , we find p, jjbpr, and Cr^ 

The loads per imit length in the thrust plane and in the plane of rota- /267 
tion as well as the torque per unit length are determined from the following 
formulas : 

^^=bCR^pcos{v-pr-^)s\gnCy; (8.9) 



The coefficients of thrust and power of the rotor are determined by numeri- 
cal integration of eqs.(8.9) and (8.11). 

For an approximate consideration of the tip losses, the loads per lonit 
length in the thrust plane are not integrated \sp to the blade tip (r = 1) but xsp 
to r = B whereby, according to another paper (Ref .2), we have 

B = \-A^ = \-± bo.it. (8.12) 

Equation (8.12) can be used when z^ s 3; at z^, = 2, the tip losses should 
be taken into account by more accurate methods. 

The coning angle and profile power coefficient are fo\md from the expressions 

B 
Y C dt — .— 

I ^ _ _ 



m 



pr- 



2. Results of Calculating the Characteristics of a Rotor 

The aerodynamic characteristics of a rotor with rectangular twisted blades 
286 



having a high-speed profile at the tip (-variant II in Table 2-10) for solidity 
ratios of the rotor cr = 0.0525; 0.069; O.O9I; 0.11 (the number of blades is, re- 
spectively, Zt = 3; 4; 5; 6), k = O.4, -^ — = 1.28*, and for two values of Mq 

are shown in Figs. 2.169 and 2.170. Such graphs are used in check calculations 
of helicopters in order to determine the rotor thrust in a hovering regime, when 
power, flight altitude, and the rotor parameters F, o, u)R are known. The se- 
quence of the calculation is as follows: Having calculated nit ^"^ ^ 



75 AT 



75Are;,ge 






(8.15) 
(8.16) 



the graphs can be used for finding the thrust coefficient t and for determining 

the rotor thrust 



























' 










- 


— 


/ 


7^ 








/ 




/• 








A 




/ 




/? ?n 






nS- 


/ 


\ 


/ 




/ 














.^A 


€ 




^ 


7^ 


y 








. 




















^y^;^ 


















/ 


v% 


/ 


e=75' 


-- 


















r 




/ 


/ 

A 


/ 

/ 


V 


7^ 




y 






n f^ 


y 


















~~ 












\ 


/ 


^ 


'/ 


r 




















































— 








— 






k 


"J 


^^ 


'7. 






— 




































r 














/, 






















t 


..J 








- 
























ie,-o: 























p. 00s 



0.0m 



0.015 



Fig. 2.169 Aerodynamic Characteristics 
of a Rotor in Hovering Regime (Mq = 
= 0.6). 



T'—^^i^Rf'^Ft. 



(8.17) 



The effect of the geometric /268 
blade characteristics is illus- 
trated ty the graph in Fig. 2.171, 
which indicates that the trape- 
zoidal blade (variant I) and the 
rectangular blade with an increased 
twist (variant VI) at t = 
= 0.12 - 0.15 require 3 - h% less 
power than a rectangular blade with 
moderate twist (variant II). Thus, 
an increase in geometric blade 
twist inproves the rotor charac- 
teristics in hovering and in 
forward flight (with the exception 
of the regime of autorotation) , 

Figure 2.172 shows the radial 
distribution of the axial v and 
tangential w conponents of induced 
velocity. The slope of v and w 
with respect to r has a different 
character for blades of the ex- 
amined shapes. For a rectangular 
blade with a geometric twist of 7° 
(variant II), v increases from the 
root to the tip of the blade; for 
a trapezoidal twisted blade and 



■55- For other values of k and y» the rotor pitch should be converted by eq.(6.23). 

287 



Z262 



t 






























- 


. 


/ 


rtf 
































OJO 
































rl^ 




























'A 


4' 


^^Y 


\^ 


^ 






























> 


r^ 


1^ 




p 


p^ 
























- 


/ 


/ 




> 




g=/r 
























>^ 


/^ 








0.15 


■ 




















/ 


/ 
























/ 


p 


nz 


r 




























/ 




/ 


'. 

































/. 


'V/ 


' 
























/. 


^. 




■ 






fiw 
















/. 


'/ 


■;/7; 


















f 




// 






































// 


/A 
































V/ 


^•75- 




















y 




















0.05 


















































y 










































































J 


'm; 


. 




























/ 










































le,=o; 

































fti7^i" 



fli'W 



«^7,f 



OT, 



Fig. 2. 170 Aerodynamic Characteristics of Rotor in 
Hovering Regime (Mq = O.?). 



o.n 



0,10 



0.05 

























- 


^ 


^ 


^ 


B 

? 


^ 






















- 






























































A 


y 






















// 


y 






























A 


/ 






























/ 


y 




















J 


p 
































/ 






















































































'-"'6 
























— 








f 























0M5 



0.010 



0.015 m^ 



Fig. 2. 171 Aerodynamic Characteristics of Rotor with 
Blades of Different Shape in Hovering Regime 
(a = O.O9I; Mo = 0.7). 



288 



for a rectangTjlar blade with increased twist, the distribution of v in the tip 
portion of the blade is close to \aniform, so that these blades have smaller in- 
duced losses. As indicated in Fig. 2. 172, the slipstream velocity w of the heli- 
copter rotor is by one order of magnitude less than the axial induced velocity v. 

For rectangular blades, the angles of attack of the sections a^ (Fig. 2. 173) 
decrease toward the blade tip, and the maximum angles of attack of the sections 
are at r= 0.3 -0.5. The trapezoidal twisted blade has a more imiform distri- 
bution of Qfp over the outside half of the blade; this angle of attack distribu- 
tion, conpared to the rectangular blade, leads to earlier attainment of critical 
angles of attack and to a marked increase in c^p at the effective blade portion. 



t/ 


Q.OIO^ 




















0.10 




- 






























(lOS 




/f 


K 
















\-w^. 





_, 












^ 




.53 




i 


eT-" 


/ 




■ 1 ' 












£r 


O.O't 


O.OC 

-\4 

O.OOZ 


} 


y 


1 






















O.Ol 





























«r 


— 


-y 


n 














/ 


N 










J 




\ 














I 


r 




[^ 


s^""*^ 


^ 




- 




^ 










^>. 1 








\, 


^ 












s 










\ 





























/270 



0,1 P.if 0,6 O.B 1.0 r 



0,1 O.if 0.6 O.S 10 r 



Fig. 2. 172 Radial Distribution of 
Axial V and Tangential w Coiiponents 
of Induced Velocity (t = O.I5). 



RLg.2.173 Radial Angle of Attack 
Distribution of Sections (t = O.I5). 



Figure 2.174 shows the dependence of profile power on the thrust coeffi- 
cient t and Mq for a blade of the variant II. As we see from the graph, the ef- 
fect of air conpressibility becomes appreciable at t > O.I5. The graph in 
Fig. 2. 175 shows the effect of the geometric blade characteristics on the profile 
power. At Mq = 0.7, the blade profile (variant III; blade without a high-speed 
profile at the tip) has the main effect while the blade shape has a smaller ef- 
fect. An increase in geometric blade twist (variant Vl) reduces the profile 
losses of the blade at large t. 

To determine the effect of air conpressibility. Fig. 2.176 shows the graph 
of Ztoioo(Mo) = mt(Mo) - mt(Mo = 0.4) for a blade of the variant II; for other 
blade shapes, this is shown in Fig. 2. 177. The conpressibility graphs permit 
converting the rotor characteristics to other Mq numbers and are also used in 
an approximate calculation of the rotor characteristics when mpj. is determined 
'by eq.(8.28). We see from Fig. 2. 177 that, for the examined profiles, the blade 
without the high-speed profile, at Mq > 0.6, shows a substantially greater inr- 
crement in mp^ than the blade with the high-speed profile at the tip. 

The values of the thrust coefficients t^p maximally permissible in view of 
the flow separation at the rotor blade (see Sect .4*7) were deteiTnined in hover- 
ing from the plot of thrust coefficient versus rotor pitch Gq . This dependence 



289 



* ' -4 'sj^''^^-^^'' 


/ '^y^^^' ^"^ ^-^ 


§f V^-^"" ^^ 


ijy /^,^~ ^ 


,„ rT1_/^ 


"' rri / 


-4~, / ' 


U-TV ^^ 


ttry 


uy 


gti 


jtrT 


\L Z 


Wl 


fi i< . 31 1 




4 


~t[ 


'~ ' ' 


Oi - ^ '^' 










tins __ . 



/271 



t 














/ 


K 


















ET/ 


• • 


/ 


. 


y^ 


" 














IVl 


/r 












1 


// 






/ 












t 


y 




m/ 










/ 




















/ f 






/ 










II 






/ 


t/.^(/ 






/ 


II 




/ 


















^ 




/ 
















// 




/ 














/ 


/ 






ft/J 






1 


1 




















1 1 
























7 
















/ 












- 












/ 










- 






1 


' 


































































- 




- 




















an^i 



















0.005 



0.010 



0,005 



0,010 mo 



Fig. 2. 174 Profile Power Coefficient as Fig. 2. 175 Profile Power Coefficient as 
a Function of Thrust Coefficient t a Fimction of Thrust Coefficient t for 

Rotors with Blades of Different Shape 
(Mo = 0.7). 



and Mq. 



6m, 



'CO 



tm 



m 



























- 


1 


1 


1 


/ 










































































1 


























i 


< 


1 


- 


/ 


























/ 




1 
























X 


/ 


/ 




/ 
























> 


7 


^ 


/ 


J 


f 


























/ 


.f 




^6. 


/ 






















/ 




/ ^ 


















/ 




/ 

y 


V 


/ 


y 






























^ 
















-< 




^ 










c 


V- 






— ■ 


■-- 


"^ 


-^ 


^i 



0.4 



0.5 



0.6 



0.7 M. 



Fig. 2. 176 Increment in Profile Power Coefficient of the 
Rotor owing to Air Conpressibility. 



290 



/272 




0.0020 



0.00J5 

0.0010 

0.0005 





0.1 Mo 



t'-O.ZO 



0.5 0.6 











m 






























1 








/ 


n 








1 




1 

1 


J 


/ 


/ 


/ 








' 


/ 


/ 






/ 


/ 




71 




'/^ 


-^ 


y 


y 



















0.1 Mo 



Fig. 2. 177 Increment in Profile Power Coefficient owing 
to Air Conpressibility, for Rotors with Blades 
of Different Shape. 



0.35 - 



0.30 



0.15 



1 


"~---^ 


~ ==~~"~'^ '"^^ 


"~" ] — -— ^ ""■•-- ^ v^ 


"~" ^~- — / — ~- ^ ^ 5 


~-~ r^-~-, ^~^i^ 


^~"?~- ~^^~--, ^^ 


~'~'~-— ^ "^~~^ 


■~--~^ 


"^« 











0.H 



0.5 



0.B 



0.7 Mo 



Fig. 2. 178 tnax S'S a Fimction of Mq, for Rotors 
with Blades of Different Shape. 



291 



is linear tp to some value of t, after which the linearity is distiirbed; the in- 
crease in t with increasing Gq decreases, after which t reaches a maximtmi t^^x* 
which is taken as t^^ when constructing the limit of separation (Figs. 2. 119 
to 2.121). 

Figure 2.178 gives the graph of t„ax ^■s a function of Mq. The diagram 
shows that, in hovering, t„ax decreases appreciably with increasing Mq. For a 
trapezoidal blade, t„ax is smaller than for a rectangular type. An increase in 
geometric blade twist will increase t^ax • The blade without a high-speed pro- 
file at the tip has a larger tj,ax » however, as soon as Mq increases the differ- 
ence in t„ax will lessen. 

According to the momentum theory and with an approximate consideration of 
tip losses, the solidity ratio does not affect the angle of attack distribution 
over the radius or the magnitude of the coefficients mpj. , Amgo , t„ax» consequent- 
ly, the graphs in Figs. 2. 173 - 2.178, constructed for o = O.09I, are valid for 
all CT. 

3. Approxijiiate Method of Detennining the Dependence /273 

of mt on t 

For vertical flight regimes of a helicopter, we can obtain an e^qjression 
for mt analogous to eq.(3.67) derived in Section 3 for flight with a horizontal 
velocity of 

I _ _ 1 ^_ _ 
/n^ = J (div-\-dqw)-{-j c^p bU^ dr~tVy. (8.18) 

In this expression, Vy < at a gain in altitude. 

Introducing the designations for the terms representing the coefficients 
of induced and profile losses of the rotor, 

I 



mind = ldtv-\-dq'w; (8.I9) 



1 _ 
mpr=-^c^pbU^dP, (8.20) 



we can represent the eD^jression for m^ in the form of 

"It = m,„d -^-mpr-tVy. ( 8 . 21) 

Let us derive the approximate expressions for the conponents of m^ in 
hovering . 

To determine mimi , let us first assume that the induced velocity v is dis- 
tributed Tiniformly over the blade radius and that w = 0. 

Multiplying the left and right sides of eq.(8.4) ly cos (|j,pr - p) and using 

292 



eq.(8.9), we obtain 

^■h-dr=dCr. (8.22) 

The average induced velocity over the disk v^y is found after integrating 
eq.(8.22) with respect to the blade radius from r = to r = B: 

^«^=§^- (8.23) 

According to eqs'.(8.22) and (8.23), at constant induced velocity the ele- 
mentary thrust coefficient is distributed linearly over the blade radius and is 
equal to 

rfCr = ?^rrf?. (8.2^) 

Consequently, 

_ To take accoimt of the nonuniformity of axial induced velocity distribution 
V and of the term dqw (power losses due to twisting of the flow passing through 
the rotor), we will introduce into eq.(8.25) the coefficient i: 



'"'"<^==i^^^"'- (8.26) 



J2B3 

The coefficient $ depends on the planform of the blade and on its geometric 
twist, on the solidity ratio, and also on the thrust coefficient. Calculations 
have shown that we can take the following average values of I : for a rectangular 
blade with a twist of 5 - 9°, a value of $ = 1.05; for a blade with a twist /274 
of 12 - 15°, ^ = 1.03. For a trapezoidal blade with taper H = 3 and twist of 
5-9°, $ = 1.03. The tip-loss coefficient B for rotors with a = 0.0525 - 0.11 
can be taken as equal to 0.98. 

Thus, for a rectangular blade with a geometric twist of 5 - 9°, m^ad is de- 
termined by the formula 



m 



tnd 



=0.56 V^e'\ (8.27) 



The profile loss coefficient of a rotor mp^ is most reliably determined by 
the graphs of mp^ = f(t, Mq) (see Figs. 2. 174 and 2.175) which were calculated 
for a rotor with similar geometric characteristics. If there are no such 
graphs, then mp^ is found from 

''P.. „ , (8.28) 



293 



The first term determines the profile losses at small Mq » while the second 
teim takes account of the increment in the profile loss coefficient owing to 
• wave drag. An estimate of Amoj, can be made on the basis of the graphs in 

Figs. 2.176 and 2.177. 

The average profile drag coefficient of 
the blade c^p is determined by the profile 



polar at the section r = 0.7* for an average 
lift coefficient c. 





























~--' 


^ 




















0.005 




















\^l^ 


0.00^ 














Tf'i' 




A 


y^ 






















Vh 


^ 














0.003 














^ 












"■ 










O.OOZ 


^ 


•> 


■ 




— 




l7,'-t 


.01 


i 




— 














•j* 




f^ 












0.001 




__ 


^- 




•^ 



































































0,10 






Ota 








o.zo 

1 






07 


-0.001 


~- 


= 


- 










^ 














— 




1 


« 


<?/7/.9 










-0.002 


* 


^ 


^ 












"~-^ 






- 




— 










^ 






i> 




-0.003 














*^ 


^^ 


<<?/ 


?7 


































> 




- 


—nnfiL 












































-0.005 




_ 


_ 








_ 






, 


_ 


_ 


u 



'yo 



-Zt. 



(8.29) 



Fig. 2. 179 Graph of the Incre- 
ment in Torque Coefficient of 
the Rotor in Ascent. 



The coefficient P which depends on the 
blade shape, its solidity, and on the coef- 
ficient t can be taken as approximately 
equal to unity for a rectangiilar blade and 
for a trapezoidal blade with taper T] = 3, 
p = 0.91 (see Table 2.5). 

During a vertical ascent or descent, 
the magnitude of the induced velocity of the 
rotor and thus also mj^^ will vary. There- 
fore, to determine mt at Vj_ f 0, it is not 
possible to add the term tVy to m^ in hover- 
ing flight without considering the variation 
in mini . The graph of the increase in 
torque coefficient of a rotor during ascent 



-0.0k 




as 



as a function of ?y and t has been con- 



0.0it V, 



Fig. 2. 180 Graph of Incre- 
ment in Rotor Pitch in 
Ascent . 



Ml 

structed for use in approximate calculations. 
This graph, shown in Fig. 2. 179, is obtained from 
results of calculations made by the momentum 
theory. The graph of the pitch increment A9o 
dtiring climb is given in Fig. 2. 180. 

The approximate ejqjressions for deter- /275 
mining Amago and AGq can be obtained from the 
following considerations. During an ascent at 
low Vy the average induced velocity, according 
to the momentum theory (Ref .2), is equal to 
(recalling that, with our adopted rule of signs, 
V„ < in ascent) 



Consequently, in a climb mt and Qq change by an amount of 



AW^c = flnd^^ - ntMc/.^ -tVy==t (Vase - %o^) ~ *■ ^y 



294 



(<-^)av '■a^. 2-0.7 » 



or 



Aeg=-4iK 



y 



The aulas o calctilated by the approximation formula is somewhat smaller than 
in Fig. 2.179. 

Thus, the final e^qjression for an approximate determination of the rotor 
torque coefficient as a function of the thrust coefficient and relative vertical 
speed has the form 

4 . Conversion_gf _Aerodyiiamic Characteristics on Variation 
in the Rot or Solidity Ratio 

To determine the aerodynamic characteristics of a rotor in hovering flight, 
the method of conversion of characteristics can be used. This method should' be 
enployed if reliable characteristics of another rotor, close in relative geo- 
metric characteristics, are available. 

The method of conversion of characteristics in flight regimes with forward 
speed, presented in Section 6, is based on a determination of the angle of at- 
tack of the rotor at which the velocity polygons, angles of attack of the sec- 
tions, and elementary force are retained for a rotor with another a in all blade 
sections, i.e., when there is similarity of regimes. In hovering, there are no 
similar regimes for rotors with different a ; therefore, the method of conversion 
of characteristics is based on the assunption that, at an identical thrust coef- 
ficient t, the induced power coefficient in conformity with eq.(8.25) is propor- 
tional to — g- and the profile power coefficient is identical. 

Thus, if the torque coefficient of the rotor of a solidity ratio a^ - m^ 
is known, then for a blade of a solidity ratio 03 the coefficient m^ for the 
same value of the thrust coefficient is determined by the formula 



m*_ = m 



t.f="''ind 






(8.31) 



The values of mp, and B are found from eqs.(8.28) and (8.12). 

At an identical value of t for rotors with different o, the angle of 7276 

295 



attack of the blade sections should be practically the same; consequently, the 
rotor pitch varies by a quantity A9o proportional to the difference of the 
average induced velocities. 



We obtain the formula for determining A9o 5 

B _ _ B 

^ ' 



Thus, the pitch of a rotor with a solidity ratio a^ is equal to 

60 --=80 4- — l/7(l/Z-)/r,). f^ s 

2 I ' 452^^ '-'^ 2 * ^> (8.33) 

5 • Determination of Optimal Aerodynamic Parameters of a 
Rotor with Consideration of the Dependence of 
Characteristics on Mq 

To select the optimal parameters of a rotor, it is convenient to use 
eq.(8.34) which correlates the rotor thrust, its diameter, and required power 

T = (33.25 Y^x\^DNf\ ( 8 .34) 

Equation (8.34) includes the relative efficiency of the rotor ^\q, charac- 
terizing the relation between the power of an ideal rotor and the actual power 
consimied by the rotor^*": 



A^ 2m, 



t 



It is obvious that, at given N, D, and A (i.e., flight altitude), the 
greatest thrust of the rotor is achieved at a maxim-urn value of Tlo; therefore, 
the designer will strive to approach a maximum TIq. Usually, the graphs of Tig 
are constructed as a function of t and a for a value of Mq corresponding to the 
average proposed values of the peripheral speed of the rotor (Fig. 2. 181). From 
this graph we select a reference point (i.e., values of t, a) with a sufficient- 
ly large TIq. We can arbitrarily select the reference point regardless of the 
rotor diameter, since an inexact agreement of the value of Mq obtained at the 
chosen t, a, D with that for which the gr^h of TIq was constructed, is considered 
permissible . 



* The power of an ideal rotor is equal to the minimum possible power losses of 
a rotor which are directly related with the generation of force, while the actual 
power is equal to the sum of all power losses. 

296 



For a more accurate calculation (with consideration of the effect of Mq on 
11 o), the use of graphs T^o = f(t, ct) constructed for different Mq is inconvenient, 
since we cannot arbitrarily select a reference point on these graphs. Actually, 
eq.(S.15) can be represented in the form 



75 N .,3 

-^3 - =^,.M„. 



(8.36) 



Consequently, at -given N, D, H the product nitoMQ should have a definite value, 
and on each curve of Tlo corresponding to definite values of a, Mq, only one 
point (one nit and, respectively, one t) satisfies eq.(8.36). Two broken curves 
are plotted in Fig. 2. 181 for two values of the product m,crM§: the smaller value 
of mtoMg pertains to the larger rotor diameter. These curves show that, at /277 
given rotor diameter and Mq, only the maximum T]o on the broken curve can be se- 
lected and that it is inpossible to realize larger values of T\q . l^on an in- 
crease in rotor diameter (lower broken curve) the maximum possible values of Tlo 
are still smaller and it is not apparent from the graph whether the rotor thrust 
determined by the product TloD increases. Therefore, the graphs of TJo, conr- 
structed as a function of t, a, and Mq (or reconstructed as a function of mtaM§, 
a, Mq), are not siii table for selecting the optimal parameters of a rotor, 
especially if the rotor diameter varies. 




Fig. 2. 181 Relative Efficiency of Rotor as a Function of 
t and a (Mq = 0.65). 



We will find a more convenient form of the graphs for selecting the optimal 
parameters of a rotor with consideration of the dependence of the characteris- 
tics on Mq . For this, we make use of eqs.(8.36) and (8.37); the latter is ob- 
tained from eqs.(8.15) and (8.17): 



297 



We can show that the ratio 



N 75 OTfMo 
t 



(8.37) 



mtMo 



on the right-hand side of eq.(8.37) is 



proportional to (TIoD)^ . Actually, from eq.(8.36) we determine the rotor dia- 
meter D: 

const -^ N 



D= f. '^ 



K,)'/2Mf' 



and the product TloD: 



., n-^ i^")' const /at ,^r-r-r/ t Vl^ 

^° — o / .1/2 ..3/2 —const VNI- — ^1 .. 

It is obvious from this expression, just as from eq.(8.37), that at 
given N and H the maximum thrust of the rotor T^^x is achieved at rotor para- 



/278 



meters at which 



mtMo 



has a maximtim. Therefore, to find the optimal rotor 



parameters we must construct graphs of the dependence of the ratio j^. 




am 

Fig. 2. 182 Ratio 



0.0002 

t 
mtMc 



0.0003 0.0004 0.0005 H!it6M; 

as a Function of mtoM^ at o = 0.091. 



298 



on 



nitaM^. These dependencies are shown at a = const in Fig. 2. 182 and at Mq = 



= const in Fig. 2. 183 • The maxiimiin values of 
these curves. 



mtMc 



lie on the envelopes of 



We see from Figs. 2. 182 and 2.183 that, at given D (i.e., ttligMq = const) and 
CT, there exists an optimal value of mJ and at given D and (uR there exists Oopt • 
If only D is given, then it follows from these graphs for a number of values of 
a and Mq that the rotor thrust increases with increasing a and decreasing wR 
(in the range of larger o than that shown in Fig. 2. 183, it may happen that, if 
only D is given,' also o^p^ and Moo^t exist). An increase in rotor diameter cor- 
responds to a decrease in mtaM| and this, as we see from the graphs, will lead 
to an increase in rotor thrust, especially on a decrease in a and Mq . 





\ 




































— 


- 












\ 

X 


^ 


^ 

■^ 


^ 


s 


^ 












• U 0.10 
< D.1Z 
' 0.14 
1 0.16 
n 0.16 
0.20 


— 






V; 










!ZS 


^ 




- 




— 






\ 


^^ 


















/ 


ft'""^ 








— 






20 


> 










-■ 


- 


— 




n\^^ 










<-'/ 










17.5 


/ 




<N 














^ 

^ 


^ 


15 


•s. 









0.OODI 0.0002 

Fig. 2. 183 Ratio 



0.0003 



0.0004 



m^M'o 



nitMo 



as a Function of 



ffltoMo at Mq = 0.65. 



All curves are closely spaced near the optim-um of a and Mq, so that any 
deviation from optimum values slightly changes the rotor thrust. For exanple. 



Fig. 2. 182 shows that, at mtoMo = 0.0003 and ct = O.O9I, we have 






19.1 for Mq « 0.55 and tgpt « 0.21, whereas with a 10^ increase in rotor 



-'opt 



rpm. 



i.e., for Mq = 0.6 and t « O.I76 we have 



mtMo 



= 18.8; consequently, the LZI2. 



299 



rotor thrust decreases by 1.6^. We find from Fig. 2. 183 that, at the same value 

of mtoMo and Mq = 0.65, we have ( — i ^ = 18.65 for o^pt = 0.08, t^pt = 

\ mtMo /opt 

= 0.165, whereas at a = O.O9I and t = 0.145, we have — = 18. 5, i.e., the 

TEtMo 

thrust diminishes by 0.8^, 

To take account of the variation in the tail rotor losses i^jon a change in 
the parameters of the main rotor, the graphs of single-rotor helicopters should 
be constructed in the form of the dependence 

— --^-^=-- fram m*°Mg (l + C V^a), 

rrifil + C y /nfO)Mo 



where the coefficient reads 



>_ ^fc/ \Dtalt' 



C 



\Dtaili 
I D \3, 






The coefficient k^i , which takes into account blanketing of the tail rotor 
by the tail boom, is taken to be equal to 1.03 - 1«06. 



300 



CHAPTER m /280 

AEHODYWMIC DESnGN OF A HELICOPTER 

Section 1. Basic Equations for Aerodynajni c Design 
of a Helicopter 

1. Aerodynami c Design Principle of a Helicopter 

The principle of aerodynamic design of a helicopter is calculating stable 
rectilinear flight regimes in a vertical plane in order to determine engine 
power, fuel consvmption, angles of attack, setting angles, and other character- 
istics of a rotor during fUght at different speeds and all possible altitudes. 
These data permit determining vertical speeds, range, and- dijration under dif- 
ferent flying conditions, and also represent necessary material for studying the 
equilibriimi conditions of moments (balancing and stability of the helicopter) 
and for stress analyses. 

Of primary interest is the determination of the performance data of a heli- 
copter, i.e., the limiting flight regimes: maximum and minim\jm horizontal flying 
speeds at all altitudes, ceiling, maximim rate of climb and range, minim-um power 
of horizontal flight, and vertical rate of descent on engine fail\ire. 

2. Equation of Motion of a Helicopter 

The flight characteristics of a helicopter are determined by solving equa- 
tions of stable rectilinear motion of the craft in a vertical plane. 

The equation expressing the sum of forces, equated to zero and directed 
along the flight path as well as along a normal to it, as shown in Fig. 3.1, has 
the form 

GsinQfi^i-Q^^^=-X- (1-1) 

GcosV^ = r, (1.2) 

where 

X and Y = conponents of the resiiltant aerodynamic forces of the rotor 
directed along the flight path and along a normal to it; at 
X < 0, the rotor creates a propulsive force while at X > it 
produces drag; 
Qp,p = parasite drag of the nonlifting parts of the helicopter; 
Qfi.f = angle of flight path of the helicopter to the horizontal. 

It follows from Fig. 3.1 and eqs.(l.l) and (1.2) that, in horizontal flight 
regimes (9f i.p = O), the Hft of the rotor balances its drag. In flight /281 
regimes along an inclined path, the propulsive force of the rotor conpensates 
the drag plus the resistance to motion formed by the weight conponent directed 

301 



II 




Fig. 3.1 Forces Acting on a Helicopter 
in Steady Rectilinear Flight. 



and the power to ^p(u)R)^aF: 



along the path, G sin 0f i.p . 

To determine the engine power 
in different flight regimes and to 
find regimes in which the majdmum 
power of the engines should be uti- 
lized, the equations of motion must 
be supplemented by an equation ex- 
pressing the condition of equality 
of the power absorbed by the rotor 
Npg t and the engine power trans- 
mitted to the rotor shaft 



N^t = Nl. 



(1.3) 



To reduce eqs.(l.l) - (1.3) to 
a dimensionless form, let us refer 
the forces to the product ^(u)R)^aF, 



1 sin Bf}^-\- c^ — = - t^; 

--Q{aR)2aF ° 



COS QfLp=ty; 



YQ(<"^)2<'^ 



75m 



1 



-tn,/.. 



■Q{a>R)3<,F 



Equation (1.4) can be then represented as 

V 



or 



iy tanBfip-^- c_^ — - = — t^; 



(1.4) 
(1.5) 
(1.6) 

(1.7) 
(1.8) 



In eq.(1.8) and in what follows, the subscript "h.f ." denotes quantities 
referring to horizontal flight. 

The given quantities in the aerodynamic calculation are as follows: /2S2 
flying weight of the helicopter G; 

geometric rotor characteristics (twist, planform), solidity ratio o and 
radius R; 

peripheral rotor speed coR; 
air density p and velocity of sound "a" at design flight altitude; 

S c S 
drag coefficient of nonlifting parts of the helicopter c^ = — ■:^ — ; 

F 

engine characteristics: power N = N(H) and hoiorly fuel consunption G^^ = 



302 



= G(M, H); 

engine power utilization factor 5 . 

In level fUght (.Qti.-o = 0)» "the number of given quantities is sufficient 
for determining the rotor lift coefficient by eq.(1.5). The problem of calcu- 
lating level-flLght regimes of a helicopter consists in deterniining, ty means of 
eq.(1.7) and for different velocities V, the required coefficient of propulsive 
force tjt , finding the values of m* from the aerodynamic rotor character- 

istics at known Mq, ty and t,. , and determining \,f from eq.(1.6). 

In maximum nonlevel flight regimes, m^ is known (N = N,,^ in c lim bing and 
N = in autorotation) : the problem amoiints to determining the values of tx and 
ty satisfying eqs.(1.5) and (1.?) from the rotor aerodynamic characteristics at 
different flying speeds at known M^ and m^; the flight-path angle 9f i.p is ob- 
tained simultaneously. 

3 . Various Methods of Determining Ae rodynamic Rotor 
Ch aracteristics and_Me thods of Aero dynamic Design 

As indicated in Subsection 2, the aerodynamic rotor characteristics, namely 
the interrelation of the four dimensionless rotor coefficients V, ty , t„ax, mt 
for a range of Mq corresponding to the rpm and flight altitudes of the heli- 
copter, should be known in the aerodynamic design. 

In certain methods of aerodynamic design, the rotor characteristics are de- 
termined by an approximate theory in order to obtain sinple formulas permitting 
a direct calculation of the helicopter performance data. Because of the approxi- 
mate nature of these calculation methods, they are rarely used at present. 

To increase the accuracy of aerodynamic calculations, it is e^^jedient to 
separate the problems of determining the aerodynamic rotor characteristics from 
those of determining the helicopter performance data. With this approach, the 
aerodynamic rotor characteristics can be found beforehand and plotted on special 
graphs. This eliminates the need for introducing sinplifications into the calcu- 
lation of aerodynamic rotor characteristics. In the Mil'-Yaroshenko method, 
presented in Section 2, the following form of graphs is adopted: The angle of 
inclination of the resultant aerodynamic force of the rotor to the normal of the 
flight path 5, the torque coefficient mt, and the angle of attack a are plotted 
as a function of the pitch cp for a series of values of the thrust coefficient t 
and the characteristic of the flight regime p, (see Fig. 2. 15). 

A further development of the Mil'-Yaroshenko method resulted in a more con- 
venient form of the graphs: dependence of the coefficient of p repulsive rotor 
force tj on m^ for various values of the lift coefficient ty at V = const and 
Mq = const. The graphs also give ciorves of constant values for the rotor angles 
of attack by means of which the latter can be defined (the angle of attack /283 
must be known for refining the parasite drag of the helicopter and for calculat- 
ing a wing-type helicopter or other conposites) . 

The graphs for the aerodynamic rotor characteristics can be plotted from 

303 



experiment or constructed from any rotor theory; the methods of determining the 
aerodynamic characteristics are presented in Sections 2, 4, 5, and 6 of Chapt.II. 

The method of aerodynamic design in which graphs of the aerodjmamic rotor 
characteristics are used, is presented in a very general fonn in Section 3. 

It was shown in Section 7 of Chapter II that the aerodynamic rotor charac- 
teristics can be determined by using the concepts of performance and propulsive 
efficiency factor of the rotor. The method of aerodynamic design based on the 
use of these concepts is described in Section 4« 



In many methods of aerodynamic design, the e^gjression 






derived in Section 3 of Chapter II, is used. 

These represent rather sinple but approximate calculation methods. One 
such method is described in Section 5. 

4. Calculation of Composite and Multirotor Craft 

For the aerodynamic design of conjiosite and multirotor craft by the methods 
described in Sections 2 and 3, we will construct graphs of the total coefficients 
of the lifting and advancing systems of a conposite craft: t^^ (or 6y) as a 

function of mty for ty = const. The total coefficients are found experimental- 
ly or can be obtained by calculation with respect to known aerodynamic charac- 
teristics of isolated elements of the lifting system of the craft. The design 
formtilas for determining the total coefficients are given below, for certain 
special cases. 

These formulas are also used in aerodynamic calculations based on the 
methods described in Sections 4 and 5, in which the Hft distribution between 
individual elements of the lifting system of the craft must be known. One of 
the possible methods of determining the lift distribution between rotors and 
wing is given in Section 4« In this case, the formulas derived below are used 
for determining the total coefficients of the lifting system of the craft. 

Single-rotor helicopter with wing . The simmaiy coefficients ty , t, , and 
mt„ are determined by the following expressions (Fig .3. 2): 

2-1 

t yj, = (^ COS ^arot - ^, sin Aa„^) -j- -^ _ x 



X {Cy^ cos Aa^ - c^^ sin Aa^ ), 

S 172 
'^, = (^xC0SAa^-l-j!ySinAa^f)4 — ^ — X 



(1.9) 



304 



X (% COS ^a^ -f Cy^ sin Aa^ ), 



'^ti=filr 



or, approximately, Tby 



'y.=^ + V 






Z28Zt 
(1.9') 



Figure 3*2 indicates that the angle between the plane of rotation of the 
rotor and the path velocity (or velocity of iindistiarbed flow), which we will 
call the angle of attack of the helicopter a^ , is equal to 

The angle of attack of the wing is correlated with the angle of attack of 
the rotor by 



a«r =-- a^ — Aa^ + e^^, = a -f Aa^,,, — Aa^ + b^. 



(1.10) 



In these e^qsressions we denote: 

ty, tx, % , a, Cy , Cj^ , (Xk = characteristics of isolated rotor and wing; 

Acf, ot = mean downwash angle in the rotor region, 
induced ty the wing; 
Aff,, = mean downwash angle in the wing region, 
induced iiy the rotor; 
S„ and ew = area and wing setting angle. 

The slipstreams of rotor and wing are denoted by the vectors V^' and V" in 

Fig .3. 2. Considering that the flight 
velocity is many times greater than 
V, the additional vertical velocities of 

interference Av, the velocities V and 
V are equal to 




V" = ]/'V2 + AT); 



'%r- 



V" = YV'-\- 



Af« 



V; 
V. 



(1.11) 



Fig .3. 2 Velocities, Angles of Attack, 
and Forces of a Single-Rotor Heli- 
copter with Wing. 



The sequence of determining the 
total coefficients at known S^ and e„ 
is as follows: _ 

For selected V, Mq, and ty , 

assign Cy and find ty from 
the first equation in the sys- 
tem (1.9'). 
From the wing characteristic 



305 



Cy = f (a„ )y find a,, • 

Determine the downwash angles Ao^ot ^^ Affw (see 'below, Subsect.5). 
Prom eq.(l.lO), determine the angle of attack of the rotor and from /285 
the characteristics of the isolated rotor, find tx and mt . 
Note that eq.(1.10) includes the angle of attack of the plane of rota- 
tion which, at cpi f^ 0, differs from the equivalent angle of attack (see 
Chapt.II, Sect. 2). 
Calculate t^^. 

After carrying out such calculations for several values of Cy , find the 

dependence of tx„ on mt for given values of V, Mq, and ty 



y * 

Perform 



these calciolations for a series of values of ty V, Mq, and then con- 

2-1 

struct graphs of the aerodynamic characteristics of the lifting system 
of the craft. 




Fig.3.3 Forces Created by Rotor and Tractor Propeller. 

It is obvious that, if the lift of the fuselage (nonlifting conponents) of 
the helicopter or a variation in its drag relative to the angles of attack must 
be taken into consideration, then the characteristics of the lifting system to- 
gether with the fuselage of the helicopter can be determined in the same se- 
quence . 

Helicopter with tractor propell ers. The additions to the total coefficients 
of the lifting system are OD^jressed by the following formulas, which are evident 
from Eig.3.3 (interference of the tractor prqjellers with other elements of the 
system is disregarded): 



A^ 



A^v 



:fc£_ 



— QO (co/?)2 3l/?2 



sin(a + e^^)5K0, 



— 0" (<*^)2 nR2 



COs(a+e^^); 



J^_ 



CO (<o/?)2 3l/?2 



(1.12) 



306 



75N. 



Affix 



'V 5 






— QO ((0/?)3 ji/?2 



In these e^^iressions Pt.p and Wt.p S'^® "the thrust and power of the tractor 



propeller, correlated Isy the ratio 75 ^t.^ 



- Pt.pV 



Nt. 



When using a cruise jet engine with a thrust of Pt.p on a helicopter, Aty 
and Atx will be determined by eqs.(l.l2), and we will have Amt = 0. 



Two-rotor helicopter of side -by-side configuration with a wing . For this /2B6 



helicopter ty^, t^^, and m^ are determined by the following e^^jressions: 

^j.=^(^ ^°^ ^^>r>t-ix sin Aa;.^^) H-^ ^X 



■jv_ 

p 



X(Cy cos^a„ —Cj^ sin Aa^), 
ix^'='^{i^ cos ^a^,t-\-ty sin Aa^,P + 



5^ V^ 



X 



X (Cj. cos Aa^ + Cy . sin Aa^ ), 



(1.13) 



Unlike eqs.(l.9), here bpi^o^ and Aa„ are the total angles of downwash in^ 
duced both by mutual interference of the rotors and interference between rotors 
and wing. 

The sequence of calculating the total coefficients is the same as for a 
single- rotor helicopter. 

Equations (1.12) are added to eqs.(l.l3) if tractor propellers are present. 

Two-rotor helicopter of fore-and-aft configuration . Disregarding downwash 
in the region of the front rotor caused by the tail rotor, we can obtain the 
following relations ( Fig .3 '4) : 



>!)" 



--iy^-\-i.iy, COS Aa„f. — t^, sin ^ar,u) « ty^ -{-ty„' 
^x^ = ix, + {tx ,Q.os Attnt, + iy, sin AOnt.) ^ ^x, + 
+ {ix, + ty,Aa„t,), 

02 = Oi — Aa;„f, + ASrete ^ 



(1.14) 



307 



In these expressions, we denote: 

Affrot ~ mean downwash angle in the region of the tail rotor, induced 
by the front rotor; 
= effective angle of advance of the tail rotor relative to the 
front rotor (with consideration of the difference in deflection 
of the automatic rotor pitch control): 



Ae 



r t^ 



A^iwfj = Ae^j/- 



-(D,x),+(D,x),. 



(1.15) 



The subscript "1" denotes characteristics of the front rotor and the /287 

subscript "2", of the tail rotor. 

The performance data of a heli- 
copter of fore-and-aft configuration 
can be uniquely determined if and only 
if the conditions of joint operation of 
front and tail rotors are known. Usual- 
ly, such a condition is the • relation 
between the thrust of the rotors deter- 
mined by longitudinal balancing of the 
helicopter. Knowing this relation for 
selected V, Mq, and ty , it is possible 

2-1 

to find tv, and t„ from the first 

equation of the system (1.14). After 
assigning ffi and calculating Aofpotp* we 




Fig. 3 .4 Velocities, Angles of 
Attack, and Forces for a Heli- 
copter of Side-by-Side Configura- 
tion. 



can determine a^ • From the characteris- 
tics of the isolated rotor, knowing V, 
0, ty, and a we then find t^ and mt for both rotors. Furthermore, t^v and mt„ 

are calculated and graphs of the aerodynamic characteristics of the lifting 

system are plotted. 



M. 



5. induction Coefficients of Two-Rotor Helicopters 
and Helicopters with a Wing 

Determination of the lift and drag of the system of lifting elements is a 
conplex problem for whose solution the induced velocity and loads per unit 
length in each section of the lifting elements should be found, with considera- 
tion of the effect of all vortices entering the system. When using high-speed 
conputers, solution of this problem is possible in certain cases. 

However, usually in aerodynamic designing, the confutation is limited to 
deteniiining the average downwash angles Aa of each of the elements of the lifting 
system. As shown in Subsection 4, the downwash angles permit finding the pro- 
jection of forces of all elements of the lifting system onto the direction of 
motion and normal to it. 

Equations (1.9) - (1.13) show that, for Aff> (i.e., for the vertical in- 
duced velocity conponent caused by other elements of the lifting system, Av is 
directed from the top downward), the drag of the craft increases by an amount 



308 



/288 




I'lg.3.5 Induced Velocity Distribution of Wing 
(Points 1, 2, 3, 1', 1' , y are above and below 
the Vortex Sheet). 



J/ 



■'.3 


' 


1 1 1 
























y=''<'Jl 
























rl K=^ 




















\\ 
























\ 




















■1,0 




^ 
























fl/^ 
























7TN 


\ 






















' 1 


<.\ 






















' 0.1 


\ 


tV 




















j / 


s 




is 


















t o-J~ 




^ 


h 
















V .•^> 




•^ 


<^ 


=^ 


5^ 












Ui^ 


p 

^ 


^ 


= 




■^ 


si 


?^=o 


SI 








1/ ^'-^ 


r" 




= ^E 




p 


— 


0,5 


Viii^ 


T 


/.ff 




1 


10 




^z 






,-3>aj ^ 








V 


~^£_^— ■ ^^ 


'^ ^ 






l^-^l-^'-; 


'2^ 


0.5^ 




>yT 


"nfr-^ t^^ 


4S 






-ZZ 




"" ^ ^ 


-' / 




"•' .--^' 


^ 


in- 








/-^ 





Fig. 3. 6 Induced Velocity Distribution along Wing Span z 
at Different Distances from Vortex Sheet y. 



309 



YAqt while, at H^a < 0, the drag of the craft decreases. 

The downwash angle of the i-th element of the lifting system induced by the 
n-th element is determined by the expression 

Aa, = ^, (1.16) 

where AVj is the vertical induced velocity conponent created by the n-th element 
at the focus of the i-th element, averaged over the area of the i-th element; 
Avj is proportional to the mean induced velocity in the plane of rotation of the 
n-th element: 

Av^ = yi,v^^ =^i — ^. (1.17) 

n ' 4B2V 

The proportionality factor n^ is called the induction coefficient. This 
depends on the mutual arrangement and dimensions of the i-th and n-'th elements 
of the lifting system. 

Let us recall how the induced velocity of the wing (or rotor) is distributed 
in space. At points downstream of the wing, the induced velocity increases and 
rather rapidly reaches double its initial value (Fig .3.5). An increase in in- 
duced rotor velocity will then take place within the rotor disk (Fig. 2. 3). 

At points upstream of the wing, the induced velocity is virtually equal to 
zero, while it decreases at points above or below the vortex sheet (points 1, 
l', 2, 2', 3, 3' in Fig. 3. 5). In cross section, the induced velocity of the /289 
wing with an elliptic circulation distribution has the form shown in Fig.3.6: 

Within the span of the wing or rotor ("i = — - — 7— < 1.0 j, the induced velocity is 

directed downward while at the periphery ("z s l.O), it is directed ijpward. 

let us determine the magnitude of the induction coefficients. 

The coefficients of mutual induction depend on the flight velocity (V) and 
on the angle of attack of the rotors; otir values of h cire averaged with respect 
to V, approximate, and applicable to all flight regimes at V s: 0.15. 

The rotors of a helicopter of side-by-side configuration, as is obvious 
from Fig.3.6, are located in the region where the induced velocities caused by 
the adjacent rotor are directed from the bottom vp. In this configiiration, the 
interference reduces the induced drag of the system. Here, the value of the 
mutual induction coefficients Hs . g was taken from B.N.Yur'yev's book (Ref.2) to 
which corrections were applied for the fact that the induced velocities at azi- 
muth t|f = 90° are greater than at azimuth ^ = 27CP . Therefore, the mutual induc- 
tion coefficients depend on the direction of rotation of the rotors: When the 
azimuth i|f = 90° is between the rotors, these coefficients are approximately 25% 
higher than the mean values obtained elsewhere (Ref.2), while they are about 25% 
lower in another direction of rotation. 

310 



*SJ 
















(. 


-o,s 






^ 


\ 














/ 




\ 


s 




1 




-ft* 




r 






\ 


\oin,l. 












\ 


-0.3 




^ 




N 




\ 










f 




\ 


\, 




V 


V. 


-o.z 










s 


V 




^v 
















-II.I 












^.^ 


n 














Wrjl 



I.S 



1.0 



Fig. 3. 7 Coefficient of Mutual Induc- 
tion of Rotors for Helicopters of 
Side-by-Side Configuration. 



1.0 



0.9 

0.8 

0.7 

0.6 
0.5 



\ 




















\ 




















\ 










t-?^^ 




\ 






ir 


. Jf[^ ' 




— N 


\ 




^Lt^^ ♦ 






\ 




















V 






















\ 






















X 


^ 






















^ 









0.1 o.z 0.3 O.it 0.5 y 



Fig. 3. 8 Coefficient of Mutual Induc- 
tion of Rotors for Helicopters of 
Coaxial and Fore-and-Aft 
Configuration. 



The graph of Jtg . g as a f iinction of the distance between the rotor axes 

z' = — ^ is shown in Fig.3«7« It is obvious that the optimum distance between 
R 

rotors, at which the least induced drag in forward flight occurs, is equal to 
z « 1.8. 

If, in helicopters of coaxial configuration, there is no vertical separa- 
tion of the rotors, then the coefficient of mutual induction k^o obviously will 
be equal to unity (Av = Vav)« When there is vertical separation, the induced 
velocity in the plane of the second rotor Av will decrease (Av < Vav ) so that 
Kgo < 1. The graph of k^o as a function of the vertical separation of the 

rotors y = -^, taken from another paper (Ref .2), is shown in Fig. 3. 8. 
R 

According to the general theory of induction, the mean induced velocity of 
the system of lifting elements does not depend on their stagger in the direc- /290 
tion of path velocity; consequently, for a helicopter of fore-and-aft configura- 
tion the mean magnitude of the additional induced velocity Av is the same as for 
a coaxial helicopter (at equal y) . Since the tail rotor does not influence the 
front rotor, we have K^ot =0; consequently, for the tail rotor located behind 



2AVa 



so that H 



r otj 



= 2>l. 



Thus, the mutual induction 



the front rotor, Av 

coefficients in helicopters of fore-and-aft configuration are also determined 
in accordance with Fig. 3. 8. 

In terms of the general induction theory, a decrease in induced drag for 



^«- In Chapter I, in eq.(3.22), we had h = 2k^ 



311 



•t*^ 





ysO 




Side-by~side 
configuration 



helicopters of side-by-side configtiration 
(hs . , > O) and an increase in induced drag 
for helicopters of fore-and-aft and co- 
axial configiirations (hoo > O) is e^qslained 
in the following manner: It is known that 
the induced drag of the system of lifting 
elements is directly proportional to the 
square of lift and inversely proportional 
to the mass of air participating in produc- 
ing lift, or to the effective cross section 
of the air stream (see Fig. 2. 56). In the 
ideal case (uniform induced velocity dis- 
tribution over the entire span), the effec- 
tive cross section of the air stream is 
equal to a circumference whose diameter is 
equal to the span of the lifting system. 



Fore-and- aft 
' confieuration 



Fig.3.9 Effective Cross Section 
of Air Stream for Helicopters of 
Side-by- Side and Fore-and-Aft 
Configurations . 



The effective stream cross section F^ 
for helicopters of side-ty-side and fore- 
and-aft configurations is given in Fig. 3 .9 • 
The sketch shows that, for the side-by- 
side helicopter, F, is greater than the 
area of the two rotors ( Kg , g < O); at 
"zs.s > 2, a gap effect appears and F^ de- 
creases. In the fore-and-aft configuration 
without vertical separation of the rotors (y = O), the effective stream cross 
section is the same as for a single rotor (.k^o =1); ^^ the presence of vertical 
separation, Fg increases (koo < !)• 

Now let us examine the interference between rotor and wing for single-rotor 
and fore-and-aft helicopters (Figs. 3. 10 and 3*11) • It is obvious that, if the 
wing of a single- rotor helicopter is very close to the rotor (y = O) and the 

spans of both rotor and wing are equal f Ivt = p - = 2.0J, then k„ « K^ot " I'O. 

l^on an increase in l„, due to the fact that the induced velocity is directed 
upward outside the rotor, the induced velocity of the rotor averaged over the 
wing span will decrease (h„ < 1»0), while the induced velocity of the wing 
averaged over the rotor area will change little. Correspondingly, ipon a de- 
crease in ly, , K„ will change little whereas K^ot will decrease. At T„ < 1.0, 
when the wing is underneath the ineffective blade sections, h„ will decrease. 
The graphs in Figs .3*10 and 3«11 are valid for a helicopter of fore-and-aft con- 
figuration, but must be taken with consideration of the mutual longitudinal dis- 
placement of both rotors and wings; for elements located aft, h is doubled, /291 
whereas for elements located forward, h decreases to zero. 

For a helicopter of side-by-side configuration, let us examine one of the 
rotors in calculating k. At "z = 2.0, half of the wing is underneath the rotor 
and half is outside the rotor (Figs.3«l2 and 3 •13); therefore, h„ and H^ot ^^® 
smaller than 0.5 (they would be equal to 0.5* if the induced velocity were equal 
to zero outside the rotor disk or wing span and were uniformly distributed with- 
in their confines). I^on a decrease in z, all larger conponents of the wing 



312 



l.O 



o.s 

















__ 
















^ 
















/ 




















/ 




















/ 






















/ 



































































































































1.0 



i.O I, 



'Rotor 



-^^ 



'if>=270' 



vwo- 



2.0 



u >z.o 



Fig .3. 10 Induction Coefficient of Rotor for a Wing on 
Helicopters of Single-Rotor and Fore-and-Aft 
Configurations • 



and rotor will be within the field of induced downward velocities, so that both 
K^ and Hyot will increase. 

The nvunerical values of h, and n^a\ plotted in the graphs (Figs.3-10-3»13) 
are given for y = 0. A decrease in k at y ^ can be determined from the graph 
of Keo = f(y) in Fig.3.8, i.e.. 



''/(!/l) = '</(!/ = 0)''co(i/l)- 



(1.18) 



Thus, in accordance with eqs.(l.l6) and (1.1?), the downwash angles are de- 
termined by the fonniilas 



Cy Cy iy o 

Aa,=x, — ==-,=0.26x, =^=0.26x, -^. 



(1.19) 



The downwash angle of the rotor induced by the wing can also be found from 
the ea^ression 



Aa„* — t-rot "7 — — '*-rot 72 77; 



aX 



ll^v^' 



(1.20) 



313 



m2 



Xrot 



i.U 






























, , 






















^ 


^ 




























/ 






























/ 
















0.S 














/ 




























t 


r 






























/ 





















































































































10 



10 3.0 L 



-Wing 
■Rotor 




l^v <z 






Fig, 3. 11 Induction Coefficient of Wing for a Rotor on 

Helicopters of Single-Rotor and Fore-and-Aft 

Configiorations 



o.s 



X 


















X 


*^ 


















V 










Oir.fl. 












' — 


—- 




■5— ' 


\ 


"V, 










































~~' ~ 


— 


CLs 


1^ 

Dirfl. 
















@i 



US 



2.0 



^te|^ 



0.3 
O.l 
0.1 











■s 


S 





^ 


_ 










- 




— 


- 


— 


" 



















uo 



1.5 



2.01 




Fig.3.12 Induction Coefficient of Rotor Fig.3.13 Induction Coefficient of 
for a Wing on Tandem Helicopter. Wing for Rotor on Tandem Helicopter. 



31ff 



Section 2. Aerodynamic Helicopter Design by the /293 

Mil^-Yaroshenko Method 

Let us examine steady regimes of rectilinear motion of a helicopter with 
low flight-path angles to the horizontal. 

Assuming the thrust of the rotor to be approximately equal to the weight 
and considering the revolutions of the rotor to be given, flight should always 
take place at a constant thrust coefficient t. In this case, the magnitude of 
the projection of the resultant onto the direction of motion can be varied only 
after having changed the angle of attack; at the same time, also the rotor pitch 
must be changed and hence the power transmitted to the rotor, so as to maintain 
balance of forces with respect to the vertical. 

The method proposed below for designing a helicopter assimies, for each pos- 
sible value of rotor pitch, that the aerodynamic rotor characteristics (thrust, 
longitudinal force, and torsion) aire known. 

1. Equations of Moti on and Design Principle 

Figure 3-14 shows the forces acting on a helicopter in steady rectilinear 
motion . 




G 

Horizontal flight 



Fig. 3. 14 Forces Acting on a Helicopter in 
Steady Rectilinear Motion. 



The equations of motion of a helicopter can be written in the form 

} (2.1) 

The angle between the direction of the resultant and the nonnal to the path 



/?sin8 + Osine/i^-|-Q^,= 0; 
;?cos8 — Gcoseyi^=0. 



reads 



8 = a+ /a;,-/ y- = a +/■«/?-' y-.. (2.2) 

315 



0.0W 



0.005 -I 




mk 



0^-t.O 



Fig .3. 15 Aerodynamic Characteristics of Rotor 
(p, = 0.15; t = 0.13; a = O.065). 





-2 
-4 
-5 
-8 
-«? 

-n 



























-%'-' 






:_i, 


1 








= 


^ 


;? 


«^ 


y 








innn_ 




„^. 










^'rf.-a ^ 


M'^ZOOOm 






^ 






— 


r- 












r ■= — 


















































~"~ 


rc;^ 






1 H=mOm 








-0.1 






•^ 






^ 




_^ 






^ 


yiLIii— • 


- 


■~, 


- 




<>'</ 


te \ 






'^ 


-— 




'\s 


-0.2 
S 








__^ 




1 ■ 




















[ ."^ 


























?3 






























^ 






























"^ 


•i 


•^ 


^=20on^ 












otrfte • 












■^ 




m 


ffffl 












-^ 


— 


— 




— 




V 


H»0 






















.^ 


-' 




•n " 


N, 












^ 


^- 


'■' 








■■ 


















^ 
















-J 







ftta 



0.20 



a3o (U 



Fig. 3 '16 Required and Disposable Characteristics 

of Helicopter. 



316 



Below, we will assume that the angles 6 and 9fi.p are small. Fiorthermore, 

owing to the smallness of -^ in flight regimes, we can assume the resultant 

force of the rotor as equal to the thrust (R = T). Then eqs.(2.1) can be re- 
written in the form 

T=Q. \ ^'^•^'' 

The angle of inclination of the forward resultant 61, . ^ required for hori- /295 
zontal flight is found from eqs.(2.3), setting 9f i.p = 0:' 

^V=-^- (2.4) 

The flight-path angle for any given regime will be determined then from 
eqs.(2.3): 



%i.p=~{^-h.f)=~l^'- 



(2.5) 



Thus, the problem consists in determining the possible angles of inclina- 
tion of the resultant 6 for each given regime. 

Figure 3.15 shows the angle 6, the angle of attack or, and the torque coef- 
ficient of the rotor m^ relative to the condition of constancy of the thrust 
coefficient t, as a function of the blade pitch cp for a specific flight regime \i,. 
The larger the setting angle cp, the more negative must be the angle of attack a 
of the entire helicopter, so as to maintain balance of forces with respect to the 
vertical, and the larger must be the angle of inclination of the forward resultant . 
The graphs in Fig. 3. 15 show that large setting angles often require a larger 
torque, i.e., a greater expenditure of power. Hence it is clear that, after de- 
termining the magnitude of the torque, it is possible - for exanple, from the 
total engine power as shown in Fig. 3. 15 - to obtain the maximum (disposable) 
pitch cp4ig for a given regime and hence the corresponding magnitudes of the 
angles of inclination of the resultant 641, and angle of attack a^is . Converse- 
ly, on assigning the value of 6 - for exanple, from the condition of horizontal 
flight - by means of eq.(2.4), it is possible to obtain the required blade set- 
ting angle cp, the torque coefficient mt, and the angle of attack of the rotor a. 

Figure 3. 16 gives the resiJ-tant values of 641s and 6^,f as a function of \i.. 
The graphs in Fig. 3. 16 are also the main graphs for the calculation, by means of 
which all necessary flight data can be determined. The intercept of the ciirves 
determines 



The vertical velocities as a function of (j, or V can be found from the 
formula 



(2.6) 



317 



where the quantity 9fi.p = -A5 is taken from the gr^h in Fig. 3 .16. 



2.. Determination of- Aerodynamic Roto r C harac teristics 

The quantities mt, 6, and cv as a function of the pitch 9 and at a given 
value of the thrust coefficient t - for exanjjle, similar to those shown in 
Eig.3.15 - which are necessary for calculation, can be determined e^qjerimentally 
or theoretically. 

Within certain limits, the Glauert-Lock theory gives results close to 
reality (Sect. 2, Chapt.Il). These limits are boiinded by a certain regime p, and 
by the magnitude of the thrust coefficient t, characterizing the value of the 
average working lift coefficient of the blade section Cy and thus determining 
the admissibility of ass'unptions made in the theory for the linear dependence 
of Cy on the angle of attack and for the possibility of adopting an average /296 

value of the coefficient of profile 



h' 

T '"1 
0^0 



0.10 



0.005 


^ 
















1 1 
















1 


















(/)=0 


■25- 


- 


- 


1 


- 


- 






\ 




^ 


-— 


— 










0^ 






J 






— 












< 


^ 


^ 


^ 


X 













. 








^ 






0. 


^ 












"= 


■— 


.^ 




-- 


r~~ 




(7i 


-N^ 






k 


<. 














" 


mt 


— 




'^•1^4^ 




^ 


N 






k, 










\ 




^^v 


r 


N 






X 


2 


\ 






N 





S 










* 


S 




-^ 
^ 




-0.005 




X 


^ 










" 


\ 












' 






^ 










\ 


^ 








^ 


■— 








-^ 


^ 


^ 


<^ 




\ 






\ 


% 












^ 




' 




;> 


y' 






\ 















L> 


■^ 


^ 


> 


^ 














\ 














^ 


^ 


■< 




















c 




0. 


'0. 


^ 


•^ 




^ 




\ 


t 




















ndfi 


^ 


^ 




























<* 






■1 


































































-aoi 


1- 














L 


_^ 




u 


_ 








U 


L 





o.w 



0.20 



0.30 



drag Cxp = const which does not depend 

a V 

on the angle of attack of the section. 

In the aerodynamic design of a 
helicopter it is convenient to define 
the flow coefficient X on the basis of 
eq.(2.50) for the thrust coefficient 
(see •Chapt.Il). For this, we make use 
of the second equation of the above 
system (2.3) which 0:153 resses the condi- 
tion that, for any rectilinear motion, 
the rotor thrust is approximately equal 
to the weight of the craft. Thus., the 
thrust coefficient in helicopter flight 
at a given rpm in all rectilinear re- 
gimes is constant, and its value is de- 
termined from the condition T = G. Then 
X, at given values of n and cp, will be 
determined from the equation 



Fig. 3.17 Auxiliary Graphs for 
Calculation of Rotor Characteris- 
tics (p. = 0.1). 



X= 



2i 



a«,B'i 



■Ht^+4)- f^-^' 



If now we substitute the value of \ 
into the ejqjression for h (2.68), m* (2.47), and o- (2.53) in Chapter II and plot 
their dependence on cp (see Fig. 3. 15), then each point of these curves will cor- 
respond to one of the possible regimes of rectilinear flight. 



To sinplify the calculations, let us plot graphs of the quantities 

ft — -^ /•-- 
h' 



*-!" ''"a. 



t 



t 



(2.9) 



3I8 



and 



ntt 



4a„ 



(1 + f^^) 



as a function of t, where cp is a parameter. 

Figiires 3.17 - 3.20 show these graphs, plotted on the assijnption that /297 
the coefficient of tip losses is B = 1 and that the mass characteristic of the 
blade is y = 5. On a change in flight altitude, y will vary in direct propor- 
tion to the variation in air density. As a consequence of a variation in y , 

also -^ — and m^ will vary, but the changes in these quantities are small for 

values (i, ^ 0.3. 



0.20[a005 




T ^i 
0.20 



0.10 



0. 



70? 
















~n 


.. 














































—" 








? 


JO 






^ 










-- 


^ 


^ 






1 




,^ 




a 















5> 


-= 








fc= 




'^ 










Sfc 




^ 


V 


— 


— 




v« 






777'- 


>» 


^ 


^ 






■v 










L c 


3 - 

■0 




^v 






N 






N 


















^ 


s 




s 






s 


-0.005 

V 






















N 




V 


\, 


























\ 






k 






> 


\%^ 


•».. 


- 


— 


— 




— 


;s 






r^ 




^ 


% 


j> 







— 




s 




— 




— 1 




= 


= 






^ 




^ 




K 


-0.010'^ 


• / 












" 








-' 










0. 


^ 




— 






^ 


5 

t 


















'NM 


_-j 


-^ 




^ 


















— 




<f 


'0.0^ 


-^ 


























" 
































- 


1015 


















1 

















V.W 



0.Z0 



0.30 



Fig. 3. 18 Auxiliary Graphs for Calcula^ Fig .3. 19 Auxiliary Graphs for Calcu- 
tion of Rotor Characteristics lation of Rotor Characteristics 
(|j, = 0.15). (iJ- = 0.2). 



The quantities m^ and 



represent conponents of the coefficients of /298 



torque and longitudinal force due only to lift and induced drag of the blades; 
the conponents of these coefficients due to the profile drag of the- sections do 
not enter into eqs.(2.9). For values of |j, within limits from 0.1 to 0.3 and for 



the usual profile surface finish, a value of c^p 
results • 



= 0.012 gives satisfactory 



319 



■■■■■■■III iiiiiiii II inn 



llllllllllllll 



llllllllllllli nil ■IIIIIIII 



3. Calculation of Flight Data 



A selection of basic parameters usually precedes the aerodynamic design. 
Let us assume the rotor diameter as given. 

Obtainment of optimum flight data in vertical regimes requires minimim loads 

on the disk area; therefore, the rotor 
^, diameter is selected as large as pos- 

sible with respect to design and weight 
considerations. Also the magnitude of 
the solidity ratio a is mostly predicated 
on design considerations. 



0.20 



0.10 



Ws 


, 


\^ 




3s 






ilJ f"'^"^ 


0,005 \ 1 


>U w=0 




iO^^^^K^y^'A 


--^^^z-^ 




--'"N--^:;^^ 


"^ x^'^^K^^"^ 


\ ^)9>> 


'^^^S'^/^K 




^-^ ^x^^s^ s^ 


_\£P V_-'^ 


^^^^^ N^ 




^Zc^^^ i: 


-0.005 V^ =^ 


^^^'^.^N N 


IK -^ y 


2- '"* "^^ ^^ 


J ^^ -^ 


^. \ 


-^^■w ^^ >^ 


^ ^ 




^ 


-wfo o^ 


. ^ ^ 


^ 












-0,0 s 





0.10 



0.20 



0.30 



Fig. 3. 20 Auxiliary Graphs for 
Calculation of Rotor Characteris- 
tics (p. = 0.3) • 



The magnitude of blade loading 
which determines the working Cy of the 
section thus depends largely on the 
rotor rpm. For a rational selection of 
the rpm it is therefore suggested to 
assign 3 or 4 values of the angular 
rotor velocity and to perform a conplete 
calculation for these. 

As regards available power trans- 
mitted to the rotor, in the case of two- 
or multirotor configurations it is 
necessary to account for the efficiency 
of transmission and for losses due to 
cooling; in the case of a single-rotor 
configuration, the power esqjended for 
driving the tail rotor must also be taken 
into account. In first approximation, 
this power can be found for hovering 
flight and is taken as unchanged in 
forward flight, which will yield smaller 
values for the performance data in /299 



forward flight than can be e^qjected in reality. 

The sequence of calculation is as follows: After assigning several values 
of (ju, a series of values of t is derived; for given values of ^x and cp, the rotor 
angle of attack a and the coefficients h and mt are determined, and eq.(2.2) is 
used for defining the corresponding values of 6 . 

The found values of 6 and mt are plotted as a fimction of cp in the form of 
graphs similar to those in Fig .3. 15, each of which is constructed for a definite 
value of |J,. Then, plotting on the y-axis the values of the available torque 
coefficient 



m^ 



'dis 






(2.10) 



the corresponding values of disposable 6413, ofiis, ^.nd cpaig are obtained. The 
320 



I l^ll I nil II I IH 



■ IIIIIHIIHMIIIIIIIII nil I llllll IIIHIIIIII nil II 



5 

3 
Z 















^y max 


























r^ 






"^ 


^. 














n=Z00rpm ^ 




\ 
















1 1 






- 


■ts. 


N 




\ 










n=Z'iO 


y 


\. 




N 
























V 


V 


\ 








-- 




















s 




\ 










• 
















\ 


N 






77=285 






'^ 


"N 










\ 


\ 






















S 









> 


A 


>( 




50 


\Ih 




700N 


\. 






\ 


f? 


















> 


> 




■ 


v\ 



Vkmthr 



Fig. 3. 21 Rate of Cllirib of Helicopter 
as a Function of Flight Velocity. 



next step is to determine the 
values of e^.f, Q?h.f> 9h.f* and 



b. f 



required for horizontal 



flight and to construct a graph of 
these values plotted against |j, 
(see Fig .3. 16 J in the same manner 
as presented atove. 

The power required for hori- 
zontal flight N^.f is found in 
terms of the torque coefficient 
mt by means of eq.(2.10). 

Having determined, by 
eq.(2.7), the values of Vy as a 
function of. V, we then construct 
the graphs shown in Fig. 3.21. 
From these graphs, we find the 



values of V, 



and V„ 



and the 



optimim rate of clLmb V,, for each flight altitude and rotor rpm. Data corre- 
sponding to other altitudes can be determined in the same manner as that given 
above; the graphs, shown in Fig.3.15, should be constructed for values of t^ 
corresponding to a certain height on the basis of the relation 



Qo 



<o 



/ f «0 •■0 



To obtain fUght data for a helicopter with respect to height above ground, 

it is also possible to use the 
following method which does not 
require constructing the graphs 
shown in Fig. 3. 16. A change to 
another altitude is characterized 
by a change in p . The graphs /300 
will remain unchanged if the value 
of the thrust coefficient is re- 
tained . Since 



Fig. 3. 22 




150 Vkm/hr 



t = 



nR2 (u/?)2 



Graph of Helicopter Flight 
Data ( n = 240 rpm) . 



then, for constancy of t, we must 
retain the equality Pooul = PhU^h • 
Hence, we determine the value of 
oJh at which t and thus also all other coefficients remain constant. The curve 
of 6i,.f = f(iJ.) in Fig. 3. 16 remains the same, since the drag Qp^^ depends on piw^ , 
and this product does not change with height. The available torque coefficient 
must be • calculated for power at an altitude, with consideration of the new value 
of angular velocity. If the power at altitude N^ = ANq, then 



321 



Ha-ving (mt )h > the described process of obtaining 6^18, cp^iB* and ot^is is 
repeated; the resultant values are plotted on graphs as shown in Pig .3. 16. 
After -determining A6, graphs of Vy = f(V) are plotted, finding ¥„ and V, 



a function of rotor rpm for various heights taking for each height n^ = Hq 



max as 
1 



Va 



These calculations must be carried out for at least three values of rotor rpm. 
Then, recording from the graphs the values Vy and V,, ^x with respect to /301 

ID & X 

altitudes for given revolutions, a graph as shown in Fig. 3. 22 is plotted, from 
which we can determine the ceiling (dynamic) and also the variation in Vj, ^x with 
altitudes, at given revolutions. 



Vc 
















-- 




5 


>J 




u 


- 




\ 


s, 


't 


\ 


\^ 










\ 


\ 


/ 








3 




\ 


s\ 


5< 


/U = 


U.I 


k 


k 


•• 






s 


n 


% 


s 


2 
















\ 


^ 


41 




















1 











He rpm 



250 



200 



— <ft=3' 



H^zom 

woo 

2000 





^V-r-t-^ TCi-J 




150 Vkm//ir 



ZOO 



Z50 



n^rpm 



Fig. 3. 23 Rotor Pitch in Autorotation Fig. 3. 24 Rotor Rpm in Autorotation 
Regime . Regime . 

The calculations for any weight can be made just as for any altitude, i.e., 
using the graphs shown in Fig. 3. 15 and changing only the value of n in conformi- 
ty with the variation in weight. 

From the condition mt = 0, the pitch cp for an autorotation regime is deter- 
mined. After constructing graphs (Fig. 3. 23) of the dependence of cp^ on the num- 
ber of revolutions of the rotor m^, for each value of m., we can find the rotor 
rpm in an autorotation regime as a -function of flying speed and for any constant 
value of cpo . The dependences n^ = f(V) for different flight altitudes are 
plotted in Fig. 3. 24. 



4. limits of Applicability of the Method 

The presented method permits analyzing the influence of numerous parameters 
322 



that determine the flight regime of a helicopter. Moreover, the degree of accu- 
racy of the calculation of performance data, based on this method, is fully de- 
termined ty the extent to which the theory underlying the calculation of rotor 
aerodynamics yields results close to reality. 

In regimes |j, -within the limits from 0.1 to 0.3 and with thrust coefficients 
smaller than the majcmum permissible, the section angles of attack are within 
the range for which the assimptions made in the theoiy are valid (cy = 3.^0,.; 
c-r, = const). 

At large values of \i and, in particular, at large blade setting angles, 
the section angles of attack in a large portion of the disk area exceed the 
critical value, and flow separation takes place. An ultimate analysis indicates 
that the theory in these regimes gives values of the longitudinal force, and 
especially of the torque, that are lower than reality, and also produces errors 
in the angle of attack. Thus, the results of the calculation by the proposed 
method should give higher values of maximum speed if this is determined in the 
region ij, > 0.3- The assimption of a uniform induced velocity distribution does 
not hold at small |j, (|j, < O.I5). In reality, the induced losses are larger in 
these regimes owing to nonuniform induced velocity distribution so that the 
calculation will give larger values of the rate of climb of the craft. 

These errors are small (of the order of 10^) for helicopters with low disk 
area loading, but markedly increase with increasing G/F, i.e., with increasing 
relative percentage of induced losses. 

An increase in the accuracy of calculation of the flight data can be 
achieved by refining the theory or by using data obtained from wind-tunnel tests. 

Section 3 • General Method of Aerod ynamic Design for 
Rotary Wing Aircraft 

In this method of calculation, just as in the Mil'-Yaroshenko method, the 
first step is to plot - on special graphs - the aerodynamic rotor characteristics. 
Then, the propulsive .force coefficient t^ is plotted against the torque coeffi- 
cient mt, for constant values of the coefficients ty, Mfi (or V), Mq (see 
Figs. 2. 142 - 2.145). To calculate a helicopter with a combined lifting system, 
the same graphs are plotted for the total coefficients; the design formulas for 
determining the total coefficients were derived in Section I.4. Thus, this 7302 
method of calculation encompasses all types of rotary wing aircraft. 

In determining the perforaiance, i.e., in solving the equations of motion, 
no sinplifying assuirptions are made and the accuracy of the calculation is de- 
termined by the accuracy of the graphs of aerodynamic characteristics of the 
lifting system and by the correctness of estimating the parasite drag of the 
helicopter and the engine power losses. We make only the assunption that the 
performance data can be calculated separately from the balance calciilation at 
some average (for a given centering of the helicopter) value of deflection of the 
automatic pitch control Kav • This leads to an error in determining the angle of 
attack of the fuselage and wing; therefore, at great differences between k and 
Hav for a helicopter with a large wing (S„/F > 0.05 - 0.0?), the assunption is 

323 



no longer valid. 

A shortcoming of this method is its relatively great e^^jenditure of time. 
Consequently, it ranges among methods of final aerodynamic design. However, 
whenever graphs of the aerxjdynamic characteristics of the lifting system are 
available, the calculation of the performance data is not excessively laborious 
and the method can be used also for preliminary calculations. 

In the calculation, aiaxLliary graphs suitable for all craft with similar 
lifting systems and equal parasite drag coefficients are constructed. By means 
of these graphs, plotted once and for all, numerous aerodynamic design calcula- 
tions of versions of a craft can be performed, including calculations for dif- 
ferent conditions of helicopter use (variations in flying weight, rotor rpm, or 
atmospheric conditions). 

1. Construction of Auxiliary Graphs for He li-copter 
Performance Data 

In this Subsection, we present a method of constnicting auxiliary graphs 
for calculating helicopter perfxjrmance data. Strictly speaking, these graphs, 
constructed for a helicopter with a specific lifting system and specific de- 
pendence Ck on oif f are applicable only to this type or to other helicopters with 
similar lifting systems and identical dependences of c'^ on at • However, the 
graphs can be used with sufficient accuracy for all helicopters of the same con- 
figuration having identical values of o, c', and other dimensionless charac- 

teristics (for exanple, ew> S„ for a helicopter with wing) and Mq not greater 
than 0.55 - 0.6, when the blade shape does not excessively influence the aero- 
dynamic rotor characteristics. Therefore, at Mq ^ 0.55 - 0.6, the aiaxiliary 
graphs are universal. Characterizing the parasite drag of the helicopter only 
by its magnitude at cif = 0, "c^, _ , it will be assumed that the increment of "Cy. 

upon a variation in c^f can be considered identical for helicopters of the same 
configuration. ¥e will disregard the lifting force of the nonlifting elements. 

For helicopters with a narrow variation range of Mq (flight at constant 
rotor rpm; dynamic ceiling less than 5000 - 6000 m) and with a maxunim Mq less 
than 0.6 - 0.65, the aioxiliary graphs are constructed for a mean value of Mq. 
For helicopters with higher Mg, determination of the performance data for the 
mean value of Mj, leads to noticeable errors, as a result of which the auxiliary 
graphs lose their universality and can be used only for one value of Mq. 

The method of enploying the graphs for determining performance data of a 
helicopter is presented in Subsection 2. 

Auxiliary graphs for required helicopter po wer. In horizontal flight /303 
of a rotary wing aircraft (9f i.p = 0), the equations of motion (1.6) and (1.7) 
take the form 



/2 

c 



^= -/...: (3.1) 



•« - ■"■h.f 



32U 



I/2QOIt;?Z(o,/?)2 y% • (3.2) 

In eqs.(3»l) and (3 •2), the index "S" means that the coefficients ty and t, 
are totaJ. coefficients of the lifting system, of a helicopter. 

If the characteristics of the lifting system are calculated -with considera- 
tion of Cx of the helicopter, then c^ in the first equation is assumed as equal 
to zero. 

Below, we will omit the index "E". For sinplicity, we will use the term 
rotor instead of lifting system and helicopter instead of rotary wing aircraft. 
The geometric rotor characteristics will be labeled by the solidity ratio o . 

It follows from eq.(3.l) that, for a given value of V in helicopters with 
identical rotors (equal a) and equal drag coefficients, the coefficient t^ has 
an identical value. Since the aerodynamic rotor characteristics, i.e., the in- 
terrelations of the quantities ty, t^ , m^ , V, and M^ are known (see Figs. 2. 142 
to 2.145), it is possible to construct auxiliary graphs valid for all flight 
conditions of a given helicopter and for all helicopters with_equal a, c^, and 
Mq, by means of which - for any value of ty as a function of V" - we can find the 
torque coefficient m^ , angle of attack of,, . f » and angle of rotor setting 9o 

required for horizontal flight. 

The sequence of constructing the auxiliary graphs for calculating horizontal 

flight regimes will_be described for the Mi-4 helicopter with rectangular metal 

blades (o = O.O63, c_, , = 0.009 with consideration of the rotor hub, or 
' (a=o) 

Cx. _ = 0.0075 without it). The calculation is made on the basis of experi- 
mental aerodynamic characteristics of a rotor with rectangular metal blades, 
a = 0.0525* When using these characteristics for the rotor of the Mi-4, the 
conversion formulas are utilized (see Sect .6, Chapt.Il). In this case, the corir- 
version is required because of differences in the rotors with respect to their 
solidity ratio, and the difference in Mq must be allowed for. No differences 
exist in profile or quality of blade manufacture, and both blade mass character- 
istic and flapping conpensator are practically identical. The parasite drag of 
the helicopter is taken without the rotor hub (the influence of the hub is taken 
into account in the experimental characteristics of the rotor) . 

Thus, conversion of the rotor characteristics is performed by the formulas: 

(3.3) 

(3.4) 

(3.5) 
(3.6) 



325 



^,=^x. 


• 


("1-')'^ . 




4fi2K2 ' 


a:=aj 




(0, — 0) ty 




4B2F2 


/Wf = 


■■mt 


i-^-^fico; 




%- 


=K 



where t^ , ofi , m* , Qq are characteristics 'of the tested rotor. 

The coefficients and angles without the subscript pertain to the Mi-4 /304 
rotor. 

The parasite drag coefficient of the Mi-4 helicopter entering eq.(3.l) was 
determined from the curve of c^ = f (a^ ) obtained from full-scale wind-tunnel 
tests of a helicopter without rotor (RLg.3.25)« If the angle of attack a at 
zero deflection of the automatic pitch control is indicated on the aerodynamic 
characteristics of the rotor, then the angle of attack of the fuselage is related 
with the angle of attack of the rotor by the approximate expression (Fig. 3-26): 



"/^a + V-A-",.- 



(3.7) 



Here, Sf is the angle of advance of the fuselage axis to the plane of rota- 
tion, and DiKav is the difference in the angles of attack of the rotor at k ;^ 
and H = 0. For the Mi-4 helicopter, Sf = 5 and the quantity DiH^v is taken to 
be e qual to -3° . 











^x 




- 


""■T 
1 


^ ^ 










121 














"-7 


— 

































— - 





















Plane 

V 


of rotation 
atX¥-U 


f 


N 


>^-D,x. 


i 


d 


s*^ 


■^ 'Automatic pi tch 
f}^ control 


^2Tr~\ 




1 


/ 



-10 



W ot) 



Fusel age axis 



Fig .3. 25 Parasite Drag Coefficient of Fig. 3. 26 For Determining the Fuselage 
Mi-4 Helicopter vs. Fuselage Angle Angle of Attack. 

of Attack. 

Equations (3 '3) - (3'7) are used in the following sequence: For the se- 
lected values of ty and V, prescribe the angle of attack of the rotor a and find 
cxf , c'x, and t^ . Using eqs.(3«3) and (3 '4) i'or deteiTnining t^ and o-i from 

the graph of the aerodynamic characteristics, check whether the values of t^ 

and QTi correspond. If not, assign a new value of a and again find t, and ofj . 

Selection of the value of a can be done rapidly in practice. After determining 
the final value of t^ , use the graphs of the aerodynamic characteristics to 

find m^ and 0q , and determine b^ from eq.(6.18) of Chapter II by means of 

the graphs in Figs. 2. 80 - 2.88. In this case, again make use of eq.(3.3) to 
find the value of t^ corresponding to a = O.O9I, for which the graphs of Ztooo 
are constructed: 



326 



t'=t.^ 



("o-^')'^ 



4B2K 



i72 



= t. 



(0.091—0.063) 4 
4-0.96V72 



The calculation is carried out in Table 3»1« 

In hovering flight, nit ^^.s determined also from the eDq^erimental curve 
(Rig.2.1!fl) with conversion to a = O.O63 by the formulas: 



tn. 






rtti 



(3.8) 



V 
a" 



(' 


.-o)^J 




4B2V2 




^-t 


r.3 


(0, — 0) ty 



4B2V/2 



77l£, 



8m, 



TABUS 3.1 
<y=0.14; 0=0.063; Afo=0.6 



/3O5 



^*-/ 



0.15 


0,20 


0.25 


0.30 


-3.5 


—5.25 


—8.0 


—11.0 


4.5 


2.75 





—3 


0.007 


0.0072 


0.0077 


0.0081 


-0.0025 


—0.00457 


—0.00765 


—0.01155 


-0.00239 


—0.00134 


—0.00086 


—0.0006 


-0.00489 


-0.00591 


—0.00851 


—0.01215 


-0.98 


—0.55 


—0.35 


—0.25 


-4.5 


-5.8 


—8.35 


—11.25 


0.0055 


0.00545 


0.00645 


0.00875 


7.6 


7.9 


9.0 


10.2 


0.00008 


0.0001 


0.00015 


0.0002 


0.00558 


0.00555 


0.0066 


0.00895 



Having made similar calculations f Or a large range of ty , we construct 
universal auxiliary graphs for determining the characteristics of the horizontal 
flight of helicopters with a = O.O63, "c^, _ = 0.0075 (without a rotor hub), 

and Mq = 0.6. These graphs are shown in Figs .3 '27 - 3 •29. 

Some sinplification in the use of the auxiliary graph of m* , shown in 

Fig. 3. 27, changing from plgrsical quantities to dimensionless and vice versa is 
possible by constructing a graph in which the ordinate does not give m* ^ but 

327 



Z206 



'"t^,^ 1 l ^z 




""'" ^ ^ V .^ 


^^^ ^ K^* 


:^ ^ s: V \% 7 


\ v^^^ ^ n- z^ 


nnm ^ \ \ ^A -^ 4-L 


aow ^ \^K^^7 ft 


^^ V V C^^ ^J-t 


\ \ nTV ^ Vl 


\ \ -^^^^^ Till 


mm<i~^ ^ V \<?, ^^-/ ■^J7'7 


00075 ^ ^-"^ ^S" /2Z_ 


^ v>>>. t^^Tn 


\ \ io\ , a/ / 


^.-aX^^^"/ --</^ 






^^^ ^^ - 


7- 


^*-V 


0W?5 1 1 




0.75 0.20 



O.W 



0.20 



Fig. 3. 27 AuxHiaiy Graph for Calcu- 
lating Horizontal FUght Regimes: 
Required_Power Coefficient as a Func- 
tion of V, ty (Mo =.0.6; a = 0.063: 



Fig. 3 .28 AiDciliary Graph for Calcu- 
lating Horizontal Flight Regimes: 
Setting Angle of Rotor (at k ='0.55; 
Y = 4*85) as a Function of V and tj 



=^(0=0) 



= 0.0075 without rotor hub). (Mq = 0.6; a = 0.063; c,.^^ .^ 

without Rotor Hub). 



= 0.0075 



0.05 0.10 0.15 0.20 025 0.30 



-5 



-10 



-h.f 



s 


'^ 










K3 


1 


1; 




^ 


^ 


■> 










\ 


;^ 


:> 


'V__ 








s 


^ 


^ 










\ 


^ 
























ty=0.22 














t 


,=azX^ 


















— 















































ty=ai;oj2;om 
^ty =0.16; 0.18 



Fig. 3. 29 Auxiliary Graph for Calculating Horizontal Flight 

Regimes: Rotor Angle of Attack as a Function of V and t™ 
(Mo = 0.6; a = 0.063; c, = 0.0075 without Rotor Hub). 



^(a=o) 



328 



the quantity +~ which is connected -with the physical quantities ty a rela- 

tion having a sijoopler form than eq.(1.6): 



i^t. 



75Ar., e 



U GaR 



(3.9) 



Such a graph is shown in Fig. 3 .30. 

In Figs .3 '27 - 3 '30 the curves are plotted to values of V permissible for 
the condition of flow separation at the rotor blades (see Figs. 2. 120 and 2.l2l). 
The curves corresponding to ty = 0.24 were obtained by extrapolation of the ex- 
perimental graphs. 

At large Mq, when the con^Dressibility effect is appreciable and the auxili- 
ary graphs become applicable only to the value of Mq for which they were con- 
structed, it is expedient to plot, for heUccpters with a turboprop engine, a 
graph for determining Mj,^, in reduced parameters: N^^ , = f(Vj. ) with the para- /307 

r 

meter G, for Mq = const (w^ = const). The reduced parameters are determined by 
the formulas: 



N, 



h.fr 



V p 



V 



Zo___ J So 

7- 75 2 



(0.379 -^)a/=-(2o,i yryy. 






V 



To 



= const, nif. 



%f' 



Vr. = ^ 1 /-^= 20. 11/7^0 -^^const^M. =const2MoV^; 
1/ / u h.f 



(3.10) 



Since, in the case of a turboprop engine, N^ determines the reduced fuel 

/— — — , it is possible to construct auxiliary 



consunption per hour G^ 



= G, 



Po 



graphs for determining G^^ and the relative fuel consunption per kilometer 

r 

q q 
-p — = -jT- in the case of helLcqpters with turboprop engines. 

AuxJli ary graph for the helicopter dynamic ceiling . From the minima of the 
ciirves of the required torque coefficients (broken curve in Fig. 3. 27, designated 
by Vh) we can construct a graph of (m^ . )Biin - f('fcy). This graph, shown in 

Fig. 3. 31, can be used for determining the minimimi required power at any flight 
altitude (at any ty) and for finding the theoretical dynamic ceiling of the heli- 
copter Hjyn , i.e., the heights at which the available power is equal to the 

minimum required power. The graph can also be used for determining the altitude 
up to which horizontal flight is possible upon failure of some of the engines /308 
of a multiengine helicopter. 

329 



0.075 



0.050 



0.025 

















-■ 






~" 




-- 




7 


"^ 














>, 


\ 








^ 


s 


\ 








— 


UJ 


^ 


^ 


^ 


N- 




'I AC' 


& 


— 


1 


^ 


^ 


W 












/ 


/ 


y// 


>l 


^ 




?> 




. 




> 


(a 


A/ 






0.76-^ 




V 


^ 


^^«?<^ 




^/ 




Xo.16 
ty=0.18 


-^ 


^ 


^i? 


^ 
^ 


^ 


— 


4 


> 






h 


O.IH 


/■■ 




































' 































0.10 



0.20 



0.3 7 



\f)min 




















- 




























0.010 




























- 
















































_, 


^ 


~ 


0.005 








^ 


-^ 








^ 


' 










































































\A. 



A 



0.10 



0.15 



0.20 



Fig. 3 .30 Auxiliary Graph for Calculating 
Horizontal FHght Regimes: Ratio — j^"''^ ■■■ as 

a Function of V and ty (Mq = 0.6; a = 0.063; 

c™, = 0.0075 without Rotor Hub). 

*(a=o) 



Fig. 3 .31 Coefficient of Minimum 
Required Power of Helicopter as 
a Function of ty (Mq = 0.6; 
a = 0.063; Cx(q^o) = 0.0075 

without Rotor Hub) . 



Auxiliary graph for maxunum rate of climb . To calculate fUght regimes of 
a helicopter in which the flight-path angle 9f i.p is not equal to zero, eqs.(1.6) 
and (1.7; must be solved for 9fi.p after determining the value of mt with respect 
to the available engine power for optimTaa rate of climb and after setting m^ = 
for gliding in autorotation of the rotor. This problem is solved either with 
the assunption of a small value of the angle 9fi.p (cos 9fi.p =1), or by suc- 
cessive approximations; however, it is more convenient to construct a universal 
aiEciliaiy graph. 

First we determined the flying speed at which the vertical speed is maximum, 
i.e., the optimum rate of climb Y^ . Calculations show that, for a helicopter, 
the optimum rate of climb practically coincides with the rate of horizontal 
fUght at which the required power is minimum. This is explained by the fact 
that the excess of rotor shaft horsepower used for climbing is maximum in this 
regime (since the available shaft horsepower of the rotor depends little on the 
flying speed) and that the propulsive efficiency of the rotor (see Sect. 7, 
Chapt.II), i.e., the efficiency of converting the excess rotor shaft horsepower 
to an excess of propulsive power creating vertica.1 speed, depends very little on 
the fljdng speed (with the exception of near-separation regimes). Therefore, 
the optimum rate of climb for all values of ty is fo-und in Fig .3. 27 from the 
curve connecting the minima of the required torque coefficients. 

It is obvious that, for all values of ty, the regime of optimum climb cor- 
responds to V = 0.15 - 0.2. 

Therefore, the aioxiliary graph for determining the vertical speed of a 



330 



helicopter is constructed for two values of V: V = 0.15 and V = 0.2; for inter- 
mediate values the vertical speed can be determined by interpolation. The 
auxiliary graph is constructed in the following sequence: 

Assign several values of 9fi.p (both positive and negative). 

From eq.(l.6), find t^ , and determine Cx as a function of a^ : 

9 



«/=%^ -W; 



Assign a mmiber of values of ty and find ty 



(3.11) 



(3.12) 



From the graph of aerodynamic rotor characteristics with respect to t^ 
and ty , determine mt for all values of ty . ® 



'9' 
Then, determine Amt (see Fig. 3 .32): 

Determine the vertical conponent of flying speed 



Z202 
(3.13) 



v.=^= 



l^sine^tp j- 



V sin %iy. 



(3.1^) 



Construct the graph of Vy = f(Amt) with the parameter ty; such a graph 
is shown in Fig. 3 .33. 

It should be noted that, because of the linearity of the aerodynamic rotor 

characteristics and because of the equidistant 
translation of the ciirves of t^ i:qDon a variation 
in a, the aixxiliaiy graph shown in Fig. 3 .33 is 
applicable for calculating helicopters with any 
c^x and CT (for Mg less than 0.6). 

If the graph is constructed for large nega- 
tive values of Am,., then the vertical rate of 
descent of the helicopter during gliding in an 
autorotation regime at a given peripheral rotor 
speed can be determined. 

To deteraiine the static ceiling of a heli- 
Fig.3.32 For Determining copter and the rate of climb in vertical ascent, 
the Increment in Power let us use the graph shown in Fig. 3 .34 which is 
Coefficient in Flight a reconstructed graph of the aerodynamic rotor 

along an Inclined Path. characteristics for V = 0. 



tx 










\ 


, 1 


x 
\ 


4m^ 


" 


h' 


X 
\ 




mf 


J 


1 








^ 


S. jr 






<! 


r^ 



2. Determination of Helicopter Performance Data 

The sequence of determining the performance data of a helicopter from 



331 




/310 



Fig. 3 •33 Atcd-liary Graph for Determining Maxim-urn 
Rate of Climb. 



^t / / / 


^y / ^^ 


^: S^t/t^X. 


,,.%jty/7. 


'T-'^tK^^y 


"•""' ^7^^,^ 


.^^^^#p i 


^^^^^v! 




„„„, .^^i^^^ 


0.005 y^>'^'^ 


gig--" 


^^ 








0.05 



a?o 



0.f5 



Pig«3»34_ Torque Coefficient as a Function of t 
and Vy for V = (Mq = 0.6; a = O.O63). 



332 



auxiliary graphs Is as follows: 

Select the design flight altitudes and calculate, for each altitude, the 
lift coefficient in horizontal flight and the available power coefficient of the 
rotor : 

*'~ 1 ~'W^^W ^ "' A ' (3-15) 

75m 



-e^:. 



'* QoF((o/?)3 (3.16) 

The design flight altitudes are selected at intervals of 1000 - 1500 m. 
The design altitudes should include the critical altitude and other salient 
points of the altitude characteristics of the engine. 

The torque coefficient m^ , angle of attack 0"^ . f , and angle of setting 
9o , required for horizontal fUght of the helicopter are foimd for calculated 
ty by interpolation from the auxiliary graphs in Figs .3 -27 - 3.30. 

Maxunum and minimum flying speeds . These are determined from the inter- 
section points of the curves oi'nit andm^ • There is no need to construct a 

special graph of m^ , and m,. , and V ^ and Vnin can be found by direct /311 

h • f d 1 s 

interpolation from Figs .3 .27 - 3 •30. If the curves of m* and m,. at large 

__ h • f d i s 

V do not intersect (at the limit of separation m^ < mt ), then the maximum 

b • f d i B 

flying speed at this altitude is not limited by the available engine power but 
by the separation of flow. 

MaxLnum vertical, rate of cTi'Tnb . This is determined from the auxiliary 
graph in Fig. 3 .33. Here, V^ and (mt , ) are found from Fig. 3. 27 for all 

calculated flight altitudes, calculating 

^'«'.ax = '"^a/^ -(/"...^Un. (3.17) 

After determining Vy from the graph in Fig .3 .33, we find 

Vy =P/o/?. (3.18) 

max 

As a typical exanple, let us determine maximxm and minimum speed, optimum 
rate of climb, and maxLmimi vertical speed of the Mi-4 helicopter with an all-ip 
weight of G = 7200 kg, tuR = I96 m/sec, and R = 10.5 m. All calculations are 
given in Table 3.2, and the results are plotted in Fig .3 .35. 

Practical and theoretical dynamic ceilings . These can be found from /312 

333 







TABLE 3.2 








H,m 





1000 


1860 


3500 


5000 


5500 


h 


0.138 


0.152 


0.166 


0.195 


0.229 


0,243 


^'"^««»V 


1430 


1500 


1550 


1315 


1380 


1300 


£(('=0) 






0.80 










0.00836 


0.00966 


0.0109 
0.84 


0.01085 


0.0134 


0.0134 


"'fe 


0.008775 


0.01015 


0.0115 


0,01145 


0.0141 


0.0141 


Vn,ax 


0.297 


0.312 


0.325 


0.305 


— 


— 


I'niax. km/hr 


210 


220 


230 


215 


— 


— 


l^mln 


0.035 


0.033 


0.03 


0,09 


0.103 


0.122 


l^mln, km/hf 


25 


23 


21 


63 


72 


85 


v« 


0.170 


0.18 


0.18 


0.20 


0.19 


0.18 


V.,. km/hr 


120 


127 


127 


141 


134 


127 


'"h/mln 


0.00537 


0.00595 


0.00665 


0.00795 


0.010 


0.0109 


im^ 


0.0034 


0.0042 


0.00485 


0.0035 


0.0041 


0.0032 


l^y 


0.024 


0.0277 


0.0292 


0.0179 


0.0179 


0.0132 


''j'max- '"/s«i^ 


4.7 


5.4 


5.7 


3.5 


3.5 


2.6 



Fig. 3 .35: The former is the altitude at which Vy =0.5 m/sec and the latter, 
the altitude at which V„ =0. From Fig .3 -35 we can determine, Xsj extrapola- 

' m ax 

tion, that the ceilings of the Mi-4 helicopter are equal to: H^y^ = 6400 m and 



H, 



dyn. 



6550 m. These data can be found without constructing a graph of Vy 



using instead the graph shown in Fig.3.31. 







/• 


'n^ 








, 


— 


- 


\ 


- 


- 


' 


- 








r 

/ 


5000 














- 




/ 




















1*000 




/ 


















/ 




Vh 


^ 








1 




3000 






/ 




















/ 






/ 


^ 












V- 


X 


r 








ZOOQ 




'/n 


n 








'ma 
















- 










- 










mo 




























i 


- 






1 






\ 




1 







m 



zoo y km/hr 



Fig. 3 .35 Flight Characteristics of 
Helicopter. 



For this, the data in Table 3«2 are 
used for plotting, in Fig.3.31, 



the curves of m^ 



dli 



tyVy 



= m* 



and m* 

Mia 

0.51 



determining, by means of eq.(3.19), the relative air density 



Tl "^diB y ouRn 

against ty (the propulsive effi- 
ciency of the rotor T| is deter- 
mined from the graphs in Sect.?, 
Chapt.Il). The values of ty at 
which these curves intersect with 
the curve of (mt j)Bin correspond 
to the theoretical and practical 
dynamic ceilings. Such construc- 
tions are performed in Fig. 3. 36 
from where we find that, at the 
practical dynamic ceiling, we have 
ty = 0.268 and, at the theoretical 
dynamic ceiling, ty = 0.274» After 



334 



II 



A = 



(3.19) 



from the standaixi atmosphere table or from the foimula 

20-«JkmL 

the ceilings are determined. In our exanple we have 

which coincides with the values obtained above. 

Static ceiling o f helicopter_ and rate of cj Limb in vertical ascent . These /313 

are found from the auxLHary graph in 
Fig. 3 .34, for which purpose the ciorve 



^t 




~~ 




" 




- 




- 


- 


















^ 








/ 












. i-,t'^ 


— 


i 


7^ 


7^ 


— 




/ 


/ 




- 




^')>\ 








- 


— 


^ 












- 




y 


0.010 


A' 








s^v.^ -^ 


L f 0.5 

^t tv — :; — 












1^ 


..»>! 






-^ 


■~ 




- 


in 




~ 



















- 












0.005 




















































1 


1 


- 






~ 












ri \\' 











of mt 



^is = f(ty) was plotted there. 



0.70 



0.15 



0.20 



0J5 t» 



Fig. 3. 36 Determination of Practical 

and Theoretical Dynamic Ceilings of 

a Helicopter (Mq = 0.6; a = 0.063; 

c". = 0.0075 without Rotor Hub). 
(a=o) 



The static ceiling of a heli- 
copter is determined under maximum 
engine operating conditions, since 
transport helicopters are generally 
not intended for prolonged hovering 
and usToally hover briefly diuring take- 
off and landing, closely above the 
field in the zone of influence of the 
air cushion. 

As a typical exan^ile, let us de- 
termine the static ceiling and vertical 
rate of ascent of a helicopter at 
takeoff power, with cuR = 212 m/sec . 
The calculations, made ty means of the 
graph shown in Fig. 3 .34, are given in 
Table 3.3^ 

From the intersection of the 
curve of m^ with the curve mt for , 

d 1 8 



V = 0, we find ty corresponding to the static ceiling, and the static ceiling 



itself: ty = 0.128; A 



^'^^^^ = 0.917; H^t = 890 m. 



0.128 



For a more conplete study of helicopter data in hovering, a graph of maxi- 
mum rotor thrust should be plotted as a function of flight altitude, for dif- 
ferent tenperature conditions t^^t with and without consideration of the ground 
effect (the latter is required for estimating the possibility of takeoff and 
landing of a helicopter in mountainous terrain). The calculation (Table 3.3) is 



335 



H, m 







h 


0.1175 





^'Vto' V 


1700 




'"'dis 


0.00784 




yy 


0.001 


( 


V, mjsec 


0.2 




t 


0.119 


( 


' taiXf k^ 


7287 





TABLE 3.3 






750 


1000 


1500 


1860 


0.126 


0.13 


0.136 


0.141 


1720 


1685 
0.8 


1600 


1560 


0.00854 


0.00858 


0.00857 


0.0086 


0.002 


—0,0005 


— 


— 


0.4 


-0.1 


— 


— 


0.1285 


0.129 


0.129 


0.130 


7320 


7170 


6820 


6630 



performed by means of the 'graph in Fig .3 .34 in terms of the ciarve for Vy = 0: 
Here, mt is determined from the available engine power, t is found from the 
graph, and the maxim-um rotor thrust T^^x is then defined. The graph of Tj,ax for 
the Mi-4 helicopter at takeoff power of the engine is shown in Pig. 3 .37. Con- 
sideration of the groTind effect on the rotor thrust is acconplished by means of 
the coefficient K^ which, for a given rotor, depends on the relative distance 



to the ground h/R. 
equal to 



Thrust with consideration of the ground effect T^ , ^ is 



(3.20) 



In Fig .3 .37, Tg,g is determined during hovering of the helicopter at a dis- 
tance of 2 m from the ground, when K^ « 1.12; this distance enables a helicopter 
of the size of the Mi-4 to take off vertically and to change to forward flight 
without touching, the ground (ground contact may take place during the takeoff 
run when the pilot deflects the helicopter and it drops slightly). 

The maximum range of horizontal flight \ t and maximum dioration of /314 
horizontal flight Tjj . f are determined by the e^^jressions : 



'Vi, 



'■h.f, 



±^ 



9inln 



[km]; 



Ohr, 



(3.21) 
(3.22) 



In these e^qsressions, we denote: 

Gf = weight of the fuel consumed in horizontal flight of the heli- 

^'^ copter; 
Gjj = hourly fuel consunption in horizontal flight of a helicopter; 
q = fuel consunption per kilometer in horizontal flight of the 
helicopter: 

^=-^- (3.23) 



336 



To determine the minim'um fuel 
consunption per kilometer, the 
minimum fuel consunption per hour, 
and. the economic and cruising 
speeds, we construct a graph of the 
fuel consunption per hour and kilo- 
meter as a function of flying 
speed. To construct the graph, we 
first use Figs. 3 .27 or 3.30 to find 
the engine power required for hori- 
zontal flight, and the engine char- 
acteristics to find the fuel con- 
sunption per hour. 

The rotor rpm at cruising and 
economic speeds should be estab- 
lished beforehand. Usually these 
are equal to the minimum permis- 
sible ipm selected by the heli- 
copter designer, on the basis of 
flight safety and design considera- 
tions; they should be combined with 
the crviising regime of the engine. 
For the Mi-4 helicopter, the peri- 
pheral rotor speed in cruising and economic regimes is equal to aoR = 180 m/sec. 
Calculation of the graph shown in Fig. 3 .38 is accomplished in Table 3.4 for an 
average gross weight of G^v « G - ^ Gf = 69OO kg. 



"^max l<S 




















8000 ' 


W=-30° 




V 


''4 










- 


— 






V 


sV 














^ 


<- 


'f 
















N. 












V 


%t 




Z — 1900 L-n 


7000 


,W=3£ 


/" 




^ 


S 


''•<n 


















X 




'*»^ 






















r^^'i 










































6000 



















woo 



mo Hm 



Fig •3-37 Maximum Rotor Thrust of Heli- 
copter in Hovering Flight. 



TABIE 3.4 



/3I5 



V 
V, km/hr 

J, kffJm 



Q 



//=1000 m\ w/?=180 m/suc; Go,=6900 kj ; <y=0.172 



0.10 
65 
0.0086 
985 
230 
3.54 



0.15 
97 
0.00725 
830 
175 
1.8 



0.20 


0.225 


0.25 


0.30 


130 


146 


162 


194 


0.00685 


0.00715 


0.0078 


0.01015 


785 


819 


893 


1162 


163 


170 


195 


308 


1.25 


M64 


1.203 


1.587 



It follows from the graph that the minimijm fuel cons-unption per hour and 
kilometer and their corresponding cmjising and economic speeds are equal to 



G, 



^r„„„=163-^ at 14,^125^ 



337 



The normal fuel load of the Mi-4 helicopter is 600 kg. From this amount, 
we must subtract the fuel consumed for starting and ground testing of the engine, 
for taxiing before takeoff, for test hovering, climbing, descending, and landing, 

and also amount of fuel needed for 
maneuvering in the air. The remainder 
of the 'unconsumed fuel is incorporated 
into the enpty weight of the helicopter 
and is disregarded in defining the 
fuel load. 

In determining the above fuel cor>- 
simption values, it is assumed that 
engine testing takes 5 niin at low speed, 
taxiing at an engine power of 0.3 of 
the rated power takes 2 min (distance 
0.3 - 0.5 km), test hovering and land- 
ing at takeoff power takes 2-3 min, 
climbing at the optimum rate takes 
place at rated power, descent proceeds 
at the most advantageous speed of Vy = 
=4-5 m/sec at 0.3 - 0.5 of the rated 
power. For transport helicopters, the 
fuel needed for navigation is assimied 



qfy/im 


^hr ka/hr 




\ 


\1 




V 


■-( 


t 


y 


/ 


/ 

/ 


A 

/ 

/ 


1.5 


zoo 








\ 










\ 














































1.0 


zoo 
















/ 


> 




- 








Cu, 








y 












^ 


■<L 


r 








y 






















Vs 
















' 




0.5 


wo 




._. 









50 



100 



150 



Vkm/hr 



Fig. 3. 38 Fuel Consimption per 
Ho-ur and Kilometer of Helicopter. 



as equal to 5% of the total fuel st^jply. 



For the Mi-4 helicopter in long-distance flight at an altitude of 1000 m, 
the sum of all fuel expenditin-es, together with the navigation svpply, amounts 
to 100 - 115 kg, i.e., to about 15 - 20^ of the total fuel load. The path /3I6 

and flying time consimied in climbing 
and descending are, respectively, equal 
to 20 km and 0.2 hr. 



Mo' const, 
(5.= const 2 
Cx=constj 




Tangent to 
origin of 
coordinate 



Thus, the Mi-4 helicopter consumes 
600 - 115 = I4S5 kg of fuel in horizontal 
flight; the maximum range and endurance 
of the helicopter are L^^f = 

485 ,,^, 485 

163 



1.16 



= 4I8 km, T, 



h.f, 



« 3 hr, while 
endurance are ] 
Tmax = 3.2 hr. 



the technical range and 
L„„ = 418 + 20 = 438 km. 



Fig .3 .39 For Determining 
and Economic Speeds 



On the assunption that the specific 
Cruising fuel consunption is independent of 

engine power and that the power utiliza- 
tion coefficient is independent of fly- 
ing speed, the regimes corresponding to 
maximum range and endurance can be detennined directly from the graphs in 
Fig. 3. 27 in the manner shown in Fig. 3. 39. 

Usually the optimum ipm in cruising and economic regimes is below that 



338 



selected by the helicopter designer. However, If the optlm-um rpm Is to be de- 
termined, calculation of Nj,,f and Ghr Is performed for several values of cuR 

h.f 

and the optlmTjm rpm Is selected from this. On the assunption that the specific 
fuel consunptlon is Independent of both engine power and engine rpm and if the 
power utilization factor does not depend on the fljring speed, the maxlmtmi range 
and duration can be determined from the following expressions: 

270O, e 1 f , 

\ ^y /mln 



''^A/niax~~ C.a^'^ 



The flying speeds and the corresponding rotor rpm (or V and ty) at which 
iuX_ and » '' - reach a minimum can be found from the graphs of these quanti- 

ties plotted on the basis of the graphs shown in Fig. 3 •27 • 

nit 
It should be noted that the quantity 1 / - ^z^ . ±q equal to the product of 

"Dy V 

helicopter performance and propulsive efficiency of the rotor K^T] [see eq.(7.lO) 
in Chapt.IIj; consequently, 

^V = ^^^^A^^lkm]. (3.240 

Minimim ver tical rate of descent . This rate, in gliding in an autorota- /317 
tion regime at a given peripheral rotor speed is deteraiined from the aioxLliary 
graph shown in Fig. 3. 33. For this, eq.(3.2) is used for calculating t„ ; for 

Amt = -(mt )nin and V^ , FLg.3-33 is used for determining Vy and then \ = 

= VyOjR. 

However, in autorotation the vertical rates of descent are determined in the 
entire range of flying speeds, both at constant rotor rpm and at constant pitch 
9o . To solve these problems, the graphs of rotor characteristics in autorota- 
tion, shown in Fig. 2. HO, are used [if necessajry, these characteristics are 
converted by eqs.C3»3) - (3 •6)]. 

For constant rotor ipm, the calculation is performed by the method of suc- 
cessive approximations. As first approximation, we use cos 9fi.p = 0.97; after 

calculating (ty \ by means of eq.(1.5), the quantities (t^ )i , a^. , o?, , c. 

' 0*11 

for a series of V are determined from the graphs of the rotor characteristics, 
and the equations of motion of the helicopter are used for finding the angle 

339 



- F2 

-^^6^.= ' , ° • (3.26) 

After repeating the calculations tintil the values of the angle Qf i.^ coin- 
cide, we find the flight-path speed and its vertical and horizontal conponents 

V = V^R; (3.27) 

Vy^VsinGfLp; (3.28) 

K^=Kcos9^l.p. (3.29) 

We note that -tan 6f i . p is equal to the inverse helicopter performance 
during gliding in autorotation: 






■ Un%fi^=-jr-. (3.30) 



In calculating the autorotation regime with a selected rotor setting (usual- 
ly 6o = 3-5°), the quantities ty , t^ , and 01^ are determined for several V 



'O ' ' ■ ~ - ''o ' "o 



from the rotor characteristics . Then, 0f 1 . p , U)R, and Mq are obtained from the 
e35)ressions : 



ta/iBflf= 



72 



0/?= A^^^^i^ 



a 

If Mq is less than 0.6, the solution is considered valid since, in this 
case, the effect of Mq on the rotor characteristics can be disregarded. 

If Mn > 0.6, the calculations must be repeated, determining ty , tjt , and /318 

C C 

Qfg for Mq obtained in the preceding approximation. The successive approximations 
are carried out rapidly and present no difficulties. After final determination 
of u)R, we determine V, Vy , and V^ by means of eqs.(3.27) - (3.29). 

As a typical exanple, let us calculate gliding in autorotation of the Mi-4 
helicopter with a gross weight of 7200 kg at an altitude of H = for coR = 
= 196 m/sec. The esqaerimental characteristics of the rotor converted to the 
solidity ratio a « O.O63 are shown in Figs .3 .40 and 3.41. 

The calciJ-ation is made in Table 3.5, and the dependence of Vy and Qti.-p on 
340 



dZ 



ai5 



0.1 

















A/ 










1/ 


















\ 






i 




'A 








/ 




















I 






7 




/ 


























i) 


/ 


- 




^/ 


















^ 


/ 
















,/r 1 


/ 














y 










', 


' 




/ 












<\ 


4^ 














J 

/ 

u 


h 


/ 


/ 














A 




















y 
















1 




/ 










.^ 


















/ 










/' 






















1 


A 


^ 






V 


/ 

















































0.01 



0.02 



0.03 



0.0't 



\ 



1223. 



Fig. 3 .40 Polars of Rotor in Autorotation Regime 
(Mo = 0.6; a = O.O63). 



"C? 






\ 






















\ 


s 






















\ 


s 






















N 


s. 
























\. 


■s^ 


























-^ 


n 























070 



0.20 



0.30 V 



Fig. 3 .41 Angle of Attack of Rotor in Autorotation 
Regime (Mq = 0.6; a = O.O63). 



700 



mVkm/hr 



-10 
V,m/s€e 















































-10 


_ 












































e/i.p 












/ 


^ 






^ 








/ 


/ 














/ 


i^ 




\ 








-20 

eh 
















\ 


Vy 




















\ 




■p 


















\, 
















\ 



Fig. 3 .42 Rate of Descent and Gliding Angle of Helicopter 
in Autorotation Regime. 



341 



V is plotted in Pig. 3 .42. This diagram indicates that the minimum vertical rate 
of descent of the helicopter is 7*2 m/sec at V = 130 km/hr, while the maximum 
gliding range, equal to 



Lgi-. 



H 






(3.31) 



is obtained for 9f i , 



ml n 



= -10°; (K^ ),,^ =5.7; V = 180 km/hr. 



TABIE 3.5 
//=0; o)/?=196m/5ec; /j,=0,138 



V 


0.15 


0.20 


0.25 


0.30 


y, km/hr 


106 


141 


176 


211 


Cvc)' 




0.1338 




(<-r,)l 


0.0315 


0.0202 


0.0166 


0.0150 


°c. 


13 


7 


4.2 


2.6 


a/=«c.+8° 


21 


15 


12.2 


10.6 


ci 




0.0067 




CxV-i 


0.00239 


0.00425 


0.00664 


0.00957 


sin e^i.^ 


—0.253 


—0.183 


—0.174 


—0.183 


e/J./' 


— 14°40' 


— 10°35' 


—10° 


— 10°35' 


cos ^jl,p 


0.967 


0.983 


0.9845 


0.983 


h. 


0,1334 


0.1356 


0.1358 


0.1356 


'-C 


0.0315 


0.0206 


0.0168 


0.0152 


sin 8^/.^ 


—0.254 


—0.183 


—0.1726 


—0.1826 


^ftp 


— 14°42' 


— 10°35' 


—10° 


— 10°30' 


Vy^V%\'a^flp,m/stc 


-7.5 


—7.2 


-8.45 


—10.7 



3. Graphs for Determining Optimum Helicopter 
Aerodynamic Parameters 



Z22Q 



The described method of aerodynamic design and the graphs of rotor charac- 
teristics used in it are convenient for a check calculation of a helicopter v/ith 
known parameters, since sufficient data are available for determining the coeffi- 



342 



cients ty and tj^ in calculating horizontal flight regimes and the coeffi- 
cient mt. in calculating climbing regimes. 

In designing a helicopter, a preliminaiy version of the parameters is se- 
lected on the basis of practical experience vjith previous models and on the 
"basis of applicable values of peripheral speed, thrust coefficient, load per 
square meter of rotor disk, etc. The next step is to refine the helicopter 
parameters. To study the effect of parameters on the performance data of a 
helicopter, special graphs should be constructed. Such graphs are necessary 
also in investigating the maximum possibilities of helicopters for inproving the 
flight characteristics. 

Calculations for aerodynamic parameter selection should be acconpanied by 
weight calculations and by investigations of the variation of parameters in a 
limited range within which the helicopter has a sufficient useful load. 

In this Subsection, a graph is described to be used for defining the rotor 
parameters ensuring the minimum required power (minimum fuel consiarption per 
hour and kilometer) at given weight, ScxS, speed, and altitude. From this 
graph, the optimum diameter, solidity ratio, and peripheral speed of the rotor 
can be determined. 

The equations for calculating horizontal flight regimes are trarieformed in 
such a manner that, in all equations, the smallest number of sought parameters 
will 'correlate dimensionless coefficients with the prescribed quantities . Equa- 
tions (3.1) and (3.2) can be reduced to the form 

'^CjS _ — <^aMg f^ 22) 

In like manner, we transform the equation for determining the required power 

N,^=±,i.Rf.Fm,^^ Q (aM,)3, ('%^) m,; 

^n,t 1 3 '"fMoMJ/ (3.34) 

2 c^s 150 —tj, 

G ^ ot 

It should be noted that the quantities Mq, Mf ^ , —, — — — are propor- 

pa pa 

tional to the reduced parameters of the helicopter: ou^ , V^ , Gp, N^^t . 

It is obvious that the required power will be lowest at a minimimi of the 

343 



1= 



TABIE 3.6 





Mo 


0.61 


0.655 




h 


0.12 


0.14 


0.16 


0.18 


0.12 


0,14 


0.16 


0,18 




tx 


-0.00853 


-0.00995 


-0.01138 


-0,0128 


—0.00853 


-0,00995 


-0.01138 


-0.0128 




Wf 


0.00825 


0.C0957 


0.01157 


0.0146 


0.0086 


0,0098 


0.01165 


0.0143 


8. 

o 


—tx 


0,03056 


0.03039 


0.03214 


0.0360 


0.0342 


0.0334 


0.0347 


0.0379 


II 




0.00558 


0.00651 


0.00744 


0.00837 


0.00643 


0.0075 


0.00857 


0.00965 




Mjc 


0.00024 


0.00033 


0.00043 


0.00055 


0.00028 


0.00038 


0.00050 


0.00063 




tx (»=0.091) 


-0.00877 


—0.01028 


—0.01181 


-0.01335 


-0.00881 


—0,01033 


-0.01188 


—0.01343 




Mf 


0.0084 


0.00972 


0.01178 


0.01483 


0.00872 


0,00993 


0.01188 


0.0146 


O 


-tx 


0.0312 


0.03085 


0.0327 


0.0366 


0.03465 


0.0338 

i 


0.0354 


0.0387 




-tjcoMl 


0.00614 


0.00715 


0.00817 


0.0092 


0.00706 


0.00824 


0.00942 


0.0106 




Mji 












! 







ratio 



mtMoMfi 



To find it, a graph in coordinates 






■n_ = f I 



t^CTM^ 






IS 



t M^ " " '■"'" 

plotted for a value of the ratio — Z¥~~ S^^^^ ^y eq'(3«33) at Mf^ = V/a 
= const. '^ 



Z222 



The sequence of constructing the graph is as follows: For the value of Mf i 
selected for the investigation, define the aerodynamic characteristics of the 
rotor in the form of a dependence, shown in Figs. 2. 105 - 2.109, for several 
values of Mq . After assigning several values to the coefficient ty and the 



G 



-, determine t^ from eq.(3.33) and find mt for each M© 



from the graphs of the aerodynamic characteristics. Then, calculate the ratio 
of the coefficients entering eqs.(3«32) and (3*34). When using the solidity 
ratio of the rotor, the quantity t, is converted by eq.(3.3) or by the formula 



tx, = tx,- 



(°i — °2) <yMo 
4fi2M5l 



In Table 3.6 a calculation is made for a flying speed of V = 275 km/hr at 

n 

an altitude of H = 1000 m (a = 336.1 m/sec, i pa^ = 6400 kg/m^ ) for -- — — = 
= 4670 kg/m^. For these data, we have ^^ 



-L.=t. 



Mfi =0.227; 
0,2272.6400 



4670 



0.071 1/„ 



TABLE 3.7 



Assigned Parameter 



0=0.091; <^R=2\2 m Isic ; 
Mo=0.63 (point a) 



Vc S 
0=34.56m; -H^L. = 0.008; 
F 
0=0.091 ( point b) 



D=34.56 m ; 



. Sc^S_ 



= 0.008; 



ioR=220 m /sic; Mo=0.655 
( point c) 



Optimum 
Parameter 


/ 150AV,/_\ 


hp 


(1>'^^\ _o.00689; 


0.0317 


6816 


\ f /opt 

D„^^=£37.43 m 






Mo^,y.=0.63; <^R,^t = 
'^=212aw/5«c- 


0,0331 


7117 






oo^ =0.095 


0.0336 


7224 



The graph for determining the optimum aerodynamic parameters is shown in 
Fig. 3 .43. Curve 1 connects the minima of the curves with identical o. From 

345 



curve 1 we find the optimuiii rotor diameter at a given rpm (Mq) and a. Curve 2 
is the envelope of the cijrves with identical ct, from which we find the optimum 
rpm at given diameter and a» Curve 3 is the envelope of the curves with identi- 
cal Mq, from which we find the optimian solidity ratio at given diameter and Mq . 



As an exanple. Table 3*7 gives the optimum parameters of a helicopter for 
G = 35,000 kg and Sc^S = 7-5 m^. 



"l , 














\ 


\ 


\, 


s 






- 


— 








^ 




f 


rc^Sf^pa 
















-j(; 




4- 




































\°^^: 


N 










> 


nnuf) 










\ 








K- 


\ 




P 


y^ 


.J 


/, 


i-a 








o,ato 








s 




^ 










S 


L 
/ 


r^ 


~ 


~- 


— 












\ 


V 




- 








^ 












\ 


•^ 










— 














\v\ 








/_ 




V 


/ 












\ 


\ 






\^^ 






i 


^T' 


A 


'J 


V 


— 


- 






— 


UJJjj 








KA 






N 


*^ 


N 
^ 










X 


"^\ 






























K 


k 

^ 








/; 


// 


n 


c 






• ty= 0.1 

0.12 

It o.n 

-A 0.16 
xty-0.18 

"Mil 












\ 


s 


V 


r. 




y 


/ 


b 






















^^ 


-/ 


^ 












0.030 
















1 


~ 


- 


— 


— 


- 


- 






































— 


^ 


















6 = 0.10 


0.0Z5 


J 

















7323 



0.0025 



0.0050 



0.0075 



0.010 ECxS 



Fig. 3 .43 Graph for Determining the Optimum Aerodynamic Para- 

G 



meters of a Helicopter (Mfi = 0.227; 



p p a. 2-1 Cj^ iD 



= 0.73). 



The above method can be used for finding the optimum aerodynamic parameters 
of a helicopter with a tractor propeller and wing; however, in this case, it is 
necessary to first determine the parameters of the prcpeller and wing (m^ , 

t„ /t„ , etc.) at which the dependence tv„ = f(mtv.) is optimum, i.e., at which 
the smallest values of t^y for all mt„ and ty„ are obtained. 



Section 4« Aerodynamic Design of a HeUcopter Based on 
Concepts of Rotor Performance and Efficiency 

The concepts of perfonnance K and propulsive efficiency T| of a rotor are 
given in Chapter II, Section 7« There graphs are presented, obtained from ex- 
periment and calculation and useful for finding the values of K and Tj. 



In this Section, we present a method of aerodynamic design of a helicopter 



346 



with the use of the concepts of performaxice K axid efficiency 71. The design 
formulas for determining required power and vertical speed of a helicopter are 
conpletely analogous to the formulas for calculating airplanes. 

This is a very sinple method of calculation, easily extended to helicopters 
of any configuration with a wing and tractor propeller or cruise jet engine. 
In a general form, it pennits making various estimate calculations in a sinple 
manner: estimating the e^^jediency of installing a wing and tractor propellers 
on a helicopter, finding the power/weight ratio M/G required for producing /32?^ 
a given maxinum speed, and determining then the amount t)y which to reduce the 
required power when reducing the parasite drag of the helicopter. 

Since the performance and efficiency yield an approximate description of 
the aerodynamic characteristics of the rotor, this method of calculation ranges 
high among the approximate methods of aerodynamic design. 

1. Helicopter Performance 

The helicopter performance in horizontal flight regime is determined ty 
eq.(7.9) of Chapter II 

ty ty Cy 

^>'^ t^-t^ ^ r^"" cx + 'cj^' ' (4.1) 

In calculations it is more convenient to use the inverse quantity, namely 
the inverse performance of the helicopter: 

i=t+V=^+V- (4-2) 

Changing to dimensional quantities, Y = G and Qp^r > we obtain 

--=— +^', (4.3) 

where the parasite drag of the helicopter is 

When using the conversion formulas for determining the performance of a 
rotor with differing parameters (see Chapt.II, Sect. 7. 6), the helicopter per- 
formance is found from the expression 

1 1 , (» — 'O^y AmprTi Qo„. 



A:^ K 4B2k2 



347 



2. Performance of Multirotor _ and CoinDOsite _Helicopters 

In the general case, the inverse performance of the craft is 

1 ItX+Q^, ^X Q^ 



K, 



^y 



a ' 



(4.6) 



where SY and Sx are the sums of lifts and drag of all lifting elements of the 
helicopter. 

Let us derive the expressions of EY and EX, for two types of helicopters - 

Single- rotor helicopter with wiuR . The lift of the helicopter lifting 
system consists of the sum of lifts of the rotor and wing 



J,v==Vr,,+y^ 



(4.7) 



¥e represent Ey in the formZ/Y= G(Y^ot + Y^ ) having designated: Y^ot = 



^'°^ and Y„ - ^" 



Sy " Ey 

In horizontal flight, we have 

'ret 'ryv ^ 1 ; 

V , ^rot ^rot . 

rot — „ : — . 



(4.8) 



y^ =■ 



'y^ ^y^ 5^ —2 I 



(4.9) 



h^ F" 



V' 



The drag is made up of the drags of the isolated rotor and wing and of pro- 
jections of the rotor and wing lifts onto the direction of motion (see Fig. 3. 11) 



^X = X,,,^X^-^Y^,t,a^, + y„ Aa^, (4.10) 

where Affro t ^-^-^ ^'^w ^^® "^^^ averaged downwash angles of the rotor and wing. 
On substituting eqs.(4.8) and (4.10) into eq.(4.6), we obtain 

Qpar 



^=^..(i+AM + ^.(^+Aa.)+^' 



(4.11) 



348 



The downwash angles are determined, as described in Section 1, by the ex- 
pressions 



^^ivt=*ret 



jiX 



t¥ 



Aa^=0.26x^^. 



(4.12) 



Two- rotor helicopter vri.th wing » After performing similar calculations, we 
find 

where the subscripts "1" and "2" denote quantities pertaining to each of the 
rotors . 

The total downwash angles due to the other two elements of the lifting 
system of the helicopter are equal to 



Aa.,r.=x„^. ^+0.26x1^', 

l^^nt, = "roi. -7- + 0.26x2 ^ , 
t,a^ =0.26x^. ^+0.26x^. "^ 



(4.1!).) 



For a helicopter of side-by-side configuration, both rotors operate under 
equal conditions (all quantities with the subscripts "1" and "2" are equal to 
each other), and k^ = Kg = Hg , g . Therefore, for a helicopter of side-by-side 
configuration we obtain the following expressions: 






Qpar 

a ' 



^^rvt — ^ivt 



aty 



nK, 



- + 0.26x^-1^; 



a/y 



Aa^=2.0.26x, ^. 



(4.15) 

/326 

(4.16) 



(4.17) 



For a helicopter of fore-and-aft configuration with a wing between the 
rotors, t?ie fixjnt rotor is virtually outside the influence of the tail rotor and 
wing, and the wing is outside the influence of the tail rotor. However, the 
induction coefficients h for the tail rotor and for the wing should be doubled 



349 






(4.18) 



The total required power of rotors of a fore-and-aft helicopter depends 
Uttle on the relation of rotor lifts. This is e^^jlained by the fact that, in 
conformity with the general theory of induced drag, this power does not depend 
on the lift distribution between individual elements of the lifting system, and 
the profile power of the rotors does not greatly depend on the lifting force 
of the rotors (in regimes not close to flow separation). Therefore, to detei^ 
mine the total required power of two rotors we can set Y^ = Yg . Actually, the 
lifts of both rotors are close in value with respect to balancing conditions of 
the helicopter. 

After setting %otj_ = %ot^ = Y^ot and K^ = K3 = K in eqs.(4.13) and (4.14), 

we find that the quantity \ can be determined by eqs.(4.l6) and (4.17), with 
the induction coefficients not doubled. The physical meaning of this expression 
is that, to determine the total power, it is possible to replace two rotors by 
one with a double Hfting force inserted between the rotors. The downwash of 
this rotor is equal to the half-sum of the downwashes of the front and tail 
rotors, i.e., equal to half of the downwash of the tail rotor. 

The sequence of calculation of helicopter performance is as follows: In a 
check calculation of a helicopter the gross weight, diameter, solidity ratio, 
rotor rpm, and parasite drag coefficient are known. After assigning the flying 
speed and altitude, find the following dimensionless coefficients: 



iyi 



— Q0/=-(a)/?)2 

M — "^ 
a 



on the basis of which, using the graphs in Section 7, Chapter II, find the rotor 
performance. Then, calculate Qp^r from eq.(4.5) and determine K^ . 

In calculating the performance of conposite helicopters, it is necessary /327 
to know the lift distribution between individual elements of the lifting system, 
i.e., Yrot , ^rot > ^w • ^o^ estimate calculations, we can assign Y„ and Cy 

for some flight regime, bearing in mind that these quantities can be obtained by 
an appropriate selection of the setting angle and the wing area. Then, using 
eqs.(4.8) and (4.9), we find Y *, ty • from the rotor and wing characteristics 

rot 

we determine K, K„. After calculating the downwash angles by eqs.(4.17), we 
350 



find Kj, . 

When the geometric characteristics and the setting angle of the wing are 
given, the following method can be used for determining rotor and wing lift in 
horizontal flight. 

The angle of pitch of a two-rotor helicopter, measured from the plane of 
rotation of the rotor (front rotor for a helicopter of fore-and-aft configura- 
tion), is determined by the following e^^jression: 



2y„f a 2Y^t 



Equation (4.19) is obtaified from the condition of equating to zero the sum 
of projections of all forces onto the direction of motion, on the assunption _ 
that the angle ^ = a^ = Ci is small, Ti = Tg, Hi = Hg = H, + TDik, H, « 0.35 VT, 
Yrot = T, X,„, = Tt? + H:Ti^ + T3 (,? + e,„J + % + H3 + Qp„ + X„ » EX = 0. 

From the angle of pitch of the helicopter, we can find the angle of attack 
of the wing 

aK'=^ + 2«r -Aa^, (4.20) 

where s„ is the setting angle of the wing relative to the plane of rotation of 
the rotor. 

— S — Q 
For known V, ty„, -r^t t„ , — Fr^^» e„, e^ot » ^ih (the desired value of 

Zi ro Li ' 

D-yH is obtained by selecting the angle of stabilizer setting), using eqs.(4.19) 
and (4.20), as well as (4.9)» (4.14;, and (4.15), all quantities enteririg these 
formulas are fo\ind by successive approximations: a„ , «?, °y„ » ^y ^j ^w* ^rot* 

We re_commend the following sequence of calculation: After assigning q?„ , 
find c„ , Y„ , K„; by means of eq.(4.15) determine 2i.ro\.> and then ty aq, . 

find «? and, from eq.(4.20), determine a^ of the second approximation. 

Two or three approximations must be performed. In this manner all quanti- 
ties for calculating K^ can be obtained. 

As an exatiple, Table 3.8 gives a calculation of the reciprocal performance 
of the Mi-4 helicopter. The initial data of this helicopter are given in Sec- 
tion 3. Performance and efficiency of the rotor were determined from the graphs 
in Figs. 2. 159 and 2.160, with conversion to the difference in solidity ratio. 
The difference in blade profiles for Mg = 0.6 can be disregarded. 

The results of calctilating -=z — for the entire range of t„ are plotted in 

Fig. 3 .4^, indicating that the inverse performance of the helicopter is minimal 
at V = 0.25 - 0.3 and at a lift coefficient close to the maximum permissible 
owing to flow separation. 

351 



The maximum performance is K^ =6.0. At small V, the reciprocal per- 
formance of a helicopter increases owing to a decrease in rotor performance and 
at large V, owing to an increase in helicopter drag. 



TABLE 3.$ 
(0.063— 0.091) ^y 



/228 



K^ Ka^ 4-0.96K2 


Y 


0.15 


0.20 


X'. 


3,5 


5.4 


T 


1.011 


0.980 


VK,, 


0.286 


0.1854 


0.063—0.091 
4-0,96K2 '" 


—0.0462 


-0.0261 


VK 


0.2398 


0.1593 


Cjc 


0.009 


0.009 


0.063<y 


0.023 


0.0408 


MK, 


0.2628 


0.2001 



^ 0.063 <y 

0.25 
6.95 
0.972 
0.144 

—0.0167 

0.1273 
0.009 

0.0638 

0.1911 



^ = 0.14 



0.30 


0.35 


8.5 


9.25 


0.964 


0.947 


0.1176 


0.108 


-0.0116 


—0.0085 


0.106 


0.0995 


0.0095 


0.01 


0.0980 


0.139 


0.204 


0.2385 



A second exanple of calctilation is that of the performance of helicopters 
of different conf igtirations : single-rotor, fore-and-aft, tandem, single- rotor 
with wing, and tandem with wing. 



The calculations were made under the following conditions: For helicopters 
without a wing, the lift coefficient of the rotor is equal to ty = 0.13, and 

rot 

for helicopters with a wing to ty^ = O.I6 and 0.32. The larger value of ty 

for helicopters with a wing corresponds to two cases: a decrease in rotor dia- 
meter when a wing is installed and a decrease in peripheral speed without a 
change in rotor diameter. The soUdity ratio of the rotor is a = O.O9I, Mq = 



= 0.65, and Mo = 0.65 - / ■ ^'j'l - = 0.587 in the latter case. The angle of wing 

setting was selected so that relief of the rotor load was equal to at least 20^ 
at M,i > 0.2. The performance and efficiency of the rotor were determined 7329 
from the graphs in Figs. 2. 159 and 2.160 and both Cy and wing performance, from 
Fig. 3 .45* The parasite drag coefficient of the single- rotor helicopter, re- 
ferred to rotor area, is equal to 0.0075 and, on a decrease in diameter, becomes 

0.0075 — ^ — = 0.00925; for two-rotor helicopters, the magnitude of Ec^S is 

twice that of the single- rotor helicopter. The wing area of the single- rotor 
helicopter, referred to rotor area, is equal to 0.0325; on a decrease in rotor 
diameter, the wing area did not change and in relative values was equal to 

352 



0.0325 -°lM_ = 0.04. 
u •Jo 



The relative wing span t^/R is equal to 0-85 and 0.95, 



respectively. For tandem helicopters, the wing area is determined by the rotor 
dimensions and is assrimed as O.I6 of the area of one rotor. The aspect ratio 
of the wings is equal to \„ = 7.2. 



0.Z5 




0.20 



0.15 



Im 


K„ 






.. 


































/ 


^ 




1 n 


20 




/ 




"■\ 






/ 






V 






f 








\, 


/ 








V 
















>. 






















,/ 


\ 




















"/ 






\ 








n c 


10 








/ 








\ 














/ 












\ 












/ 
















\ 








/ 
























A 






















/ 


/ 

























10 



zn oc„ 



Fig.3.Z|4 Reciprocal Performance of Fig .3 .45 lift Coefficient and Wing 
Helicopter as a Fimction of lift Coef- Performance as a Function of Angle 
ficient and Relative Flying Speed. of Attack. 

Calciilation of the performance of helicopters without a wing is made in 
Table 3.9, while the performance of helicopters with a wing in a version with a 
decreased rotor diameter is given in Table 3.10. 



JL 
0.3 



0.2 



0.1 











T 


n 




1 












- 






\ 


\ 


















\ 


vi 


7" 


\ 


,1 












- 






1 ' 


\} 




\, 


H 






y 






— 


\ 

-A 


v5 


^ 

v; 


i 


V. 


— 




4 




^ 


i^ 


% 


^ 


■^z 


— 


': 


■-" 




^ 


















1 













0.1 



0.1 Mfi 



Legend : 

Single-rotor configuration; 

Side-by-side config-uration; 

Fore-and-aft configuration; 

wo/w, helicopter without wing; 

I, helicopter with wing and 
reduced diameter; 

II, helicopter with wing and 
reduced peripheral speed. 



Fig. 3 .46 Reciprocal Performance of 
Helicopters of Different Configura- 
tions with and without a Wing as 
a Function of Mji . 



353 



I I 



TABLE 3.9 
HELICOPTERS WITHDUT WING 





V 


0.15 


0.20 




K 


3.5 


5.2 




1 


1.00 


0.977 




'CxV^ 


0.0142 


0.0253 




''yE 






Single- rotor configuration 


1 


0.286 


0.1925 




l^rot 




1 


0.300 


0.2178 




Xh 






Fore-and-aft configuration 


^rot 


0.143 


0.0962 


Xco = 0.65 
^,^ = 0.26 


K 

ant. 


0.181 


0.102 




)'™^J^ +AawfJ 


0.2335 


0.1472 




1 


0.3907 


0.2687 




_ '<h 


—0.0558 


—0.0314 - 


Side- by- side configuration 


^p.rot 


x„ = -0.4 
'yj,=0.26 


+ AaTOf 


0.2302 


0.1611 




1 


0.2444 


0.1864 



0.30 
7.85 
0.962 

0.057 



0.127 
0.184 

0.0635 
0.0454 

0.0862 

0.2092 

-0.01395 
0.11305 

0.17005 



0.40 
8.42 
0.936 

0.1014 



0.119 
0.220 

0.0595 
0.0255 

0,0722 

0.2373 

-0.00785 
0.11075 

0.21215 



The results of the calculations are plotted in RLg.3.46, which shows that, 
in the entire speed range, the reciprocal performance of the helicopter of side- 
by-side configuration has a lower value and that of the fore-and-aft configura- 
tion, a higher value. _The maximtim performance is equal to: 6 for a side- /33I 
by-side helicopter at V = 0.27; 5*5 for a single- rotor helicopter at V = 0.29; 
4.8 for a fore-and-aft helicopter at V =0.3. At M,i = 0.26 (V = 0.4), the per- 
formance of the helicopters is, respectively, equal to: 4*7; 4*55; and 4'23.. 



The wing, relieving 20 - 30% of the rotor load at high flying speeds, 
changes the helicopter performance in the following manner: If, on installation 
of a wing, the rotor diameter was decreased, the helicopter performance increases 
very little (ciorve I). If the rotor diameter was not decreased but its ipm was 
raised (curve II), the maximTjn performance of the helicopter increases by 
0.5 - 0.9 (by 10 - 13%) and, at maximum speed (Mf^ = 0.26), increases by 0.4 
( approximately ^%) . Calculations showed that if, on installing a wing, the 
rotor parameters are not changed so that the rotor at high speeds has a very low 



354 



TABLE 3.10 



'^s 



Single-Rotor Helicopter with ^ing 

= 0.16; c^ = 0.00925; iMo = 0.65; 5»v = 0,04; 7*v = 0.95; x,,f = 0.2; 
%H, =0.69; £„, =21.97° 



Tandem Helicopter with Wing 

<ys = 0.32; c; = 0.0185; Mo=0.65;"S^=07l6;r=7^ = 
= 1.9; «£4=— 0.4; »rrf = 0.12; x^ =0.4; i^ = 16.3° 



AoC|y, radiui 



A a^t- (from »ing) 
Act/V^ (from interference) 



Krot 
1 



+ Aa, 









0.15 
—0.45 

12.36 
0.795 

19 

0.112 
0.049 
0.951 
0.152 
3.45 
0.0065 

0.0065 
0.2963 

0.1644 

0.282 

0.00809 
0.0143 



0.20 

-2.48 

14.63 

0.915 
15.8 

0.0595 

0.10 

0.90 

0.144 

5.3 

0.0075 

0.0075 
0.1961 

0.1228 
0.176 

O.0123 

0.0254 
0.214 



0.30 


0.40 


-7.10 


-12.66 


12.7 


8.23 


0.810 


0.545 


18.4 


23.8 


0.0235 


0.0126 


0.200 


0.24 


0.80 


0.76 


0.128 


0.122 


7.8 


8.05 


0.0066 


0.0045 



0,0066 
0.1348 

0.0778 
0.1078 

0.0156 

0.0571 
0.180 



0.0045 
0.1287 

0.0546 
0.0979 

0,031 

0.1015 
0.212 



0.15 
-0.6 
8.4 
0.548 
22.5 
0.1276 
0.068 
0.932 
0.149 
3.5 

0.00145 
-0.0640 
-0.06255 

0.2274 

0.172 

0.212 

0.0117 

0.0143 
0.2.38 



0.20 
-2.9 
9.7 

0.63 
21.8 
0.066 
0.139 
0.861 
0.1376 
5.23 
0.00167 
—0.0332 
-0.03153 

0.1595 
0,1119 
0.1373 

0.0156 

0.0254 
0.1813 



0.30 
—7.62 
7,22 
0.48 
22.8 
0.0262 
0.237 
0.763 
0.122 
7.6 

0.00126 
-0.0131 
—0.01184 

0.1198 
0.0701 
0.0915 

0.0166 

0.0571 
0.1652 



0.40 
-12.5 
2.95 
0.22 
15.9 
0.0155 
0.195 
0.805 
0.1286 
8.3 

0.00058 
-0.00777 
-0.00719 

0.1134 

• 0.0783 

0.0913 

0.0153 

0.1015 
0.208! 






thrust coefficient, then installation of a wing will not result in a decrease in 
required power. 

It should be noted that an increase in thrust coefficient ty„ for a heli- 
copter with a wing leads to a_decrease in its dynamic ceiling. This is so since, 
at the optimum rate of climb Vo ~ 0.2, the wing only insignificantly relieves 

the rotor load, and ty acquires the maximum 
permissible (in view of flow separation) value 
at a lower altitude. Furthermore, at large 



0.2 



0.1 











r~ 




































^ 


<' 


■^. 
























/ 

f 






\ 




/ 
I 
















/ 
1 




/' 


• 


>^ 


















1 


/ 


4 


/ 




\ 
\ 














/ 






f 






















/ 


/ 


^ 






















/i 


/ 


/^ 


/ 






















/ 


/ 


V 
























'/ 


// 
























a 

I 


!/, 


? 
























// 


























('■ 





































































flow separation at the rotor may occur at 

is still small. To 



low flying speeds when ty 

reduce t„ at these altitudes, a 5 - 8/ 

'rot » ' / 

crease in rotor rpm can be advantageous. 



m- 



It follows from Table 3.10 that the wing 
performance with consideration of downwash by 

1 



the rotor 



(^ 



K, 



+ ba„j decreases by 



0.1 



0.1 



^fi 



Fig. 3. 47 Relative Lift of 
Wing for Helicopters of 

Different Conf igurat ions . 
Legend : 
Single-rotor con- 
figuration; 
Side-by-side con- 
figuration; 
Helicopter with wing 
and decreased dia- 
meter; 
II - Helicopter with wing 
and decreased peri- 
pheral speed. 



I - 



several units at high flying speed and even 
more at low speed. The wing, producing down- 
wash near the rotor, somewhat reduces its per- 
formance. This explains the slight change in 
helicopter performance when a wing is installed. 

On a helicopter without a tractor pro- /332 
peller, a wing without a f jjxed angle of setting 
has a maximum angle of attack «» in horizontal 
flight at V = 0.2 - 0.15. At smaller V, this 
angle decreases owing to an increase in down- 
wash from the rotor; at larger values, it de- 
creases due to an increase in pitch angle of 
the helicopter. Therefore, when a wing has a 
small area and large angles of attack, its lift 
increases at high speed despite a decrease in 
a^ , but insignificantly (Fig. 3. 47, single- 



rotor configuration). Conversely, if the wing 
has a large area and small a„ ( side-by-side 
configuration), then at large speeds Cy 

markedly decreases and the lift becomes less than at average speeds. 

Thus, on helicopters without a tractor propeller or other propeller, the 
wing should have a small area and large a„ or be provided with mechanization for 
controlling the amount of Cy . 

In a climbing regime, the angle of attack of the wing decreases, while it 
increases in gliding. At a fixed angle of wing setting in an autorotation 
regime, flow separation from the wing is inevitable, which can be tolerated in 
the presence of a small wing lift ( small wing area and reduction in Cy by 



356 



mechanization of the vrlng) . 



3. Det emanation ofH ellcopter Flight Data 



If both helicopter performance and rotor efficiency are known, the required 
power of a helicopter is determined by the expression (see Sect. 7, CHiapt.Il): 



^V 



OVlmlsec] 1 (4.21) 

755 Ki,r\ 



°^ ^ _ aVlkmlhr] J_ 

2705 /C.r,- (^^22) 

The sequence of calculation for helicopters of various configurations is 
described in Subsection 2. The rotor efficiency is determined from the graphs 
given in Section 7> Chapter II. Consequently, on assigning the flying speed /333 
and altitude, a graph of the required power of the helicopter can be plotted. 
In hovering flight, the required power is determined from aerodynamic charac- 
teristics of the rotor in a hovering regime: Nh.f is calculated at all flight 
altitudes under the condition T = G. 

The maximum and minimum flying speeds are determined from the points of 
intersection of the cvirves of required and disposable power. At all flight 
altitudes we must find the maximum permissible speed Vpg, with respect to flow 
separation conditions; if V„ ax > ^part then the flying speed of the helicopter 
is limited by the value of Vpe , . 

Having plotted the cvirves of required power and knowing the engine charac- 
teristics with respect to fuel consumption, the fuel consumption of the heli- 
copter per hour and kilometer can be plotted as a function of flying speed (see 
Fig. 3. 38) and, as described in Section 3, the maximum range and endurance, 
cruising and economic flying speeds can be determined. 

If the helicopter flight path is inclined, the propulsive force of the 
rotor should balance the projection of helicopter weight onto the direction of 

flight, which is equal to G sin Qfi.p or G ^= — (see Pig. 3.1). Therefore, the 

expression for engine power takes the form 

^=^^(T+^+~J-^*-^ + 755x1 • (4.23) 

It follows from eq.(4»23) that the maximum rate of climb of a helicopter is 
determined by the formula 

"max Q 

357 






^.fV 



2000 



1500 




WOO 



ZOO V kmjhr 



Fig. 3.4s Required and Available Horsepower 
of Helicopter. 



ams 



0.0005 




0M025 



Fig. 3. 49 Ratio Nh.f/Ga for Helicopters of 
Different Configurations with and without 
Wing, as a Function of Mfj . 
Legend: 

Single-rotor configuration; 

Side-by-side configuration; 

Fore-and-aft configuration; 

wo/w - Helicopter without wing; 

I - Helicopter with wing and decreased 
diameter; 
II - Helicopter with wing and decreased 
peripheral speed. 






The optimum rate of climb and minimum power consumption ^.f^^ are found 
from the graph of required powers. After determining Vy at all altitudes and 
constructing the graph Vy = f(H), the dynamic celling of the helicopter is 
determined by graphical means (see Fig. 3. 35). 

From eq.(4.24) for \is = 0, the minimum rate of descent of a helicopter 
in an autorotation regime of the rotor (Vy )Bin is derived. 

To determine the angles of attack a and angles of setting 9o of the rotor 
it is necessary to calculate the coefficient of propulsive force of the rotor 

Qpar+G~ 

^'~ r (4.25) 

and, knowing ty and t^ , the angles o* and Go must be determined from the graphs 
(see Figs. 2.63 - 2.70 and 2.105 - 2.109) or from eq.(3.95) given in Section 3, 
Chapter II. 

As an example, let us carry out an aerodynamic calculation of the Ml-4 
helicopter with rectangular metal blades. The graph of helicopter performance 
is given in Fig. 3. 44- The graph of the required and disposable powers is shown 
in Fig. 3. 48 for six flight altitudes. 

At V = 0, Nh.f is determined by the expression: 

where mt is found from the graphs shown in Fig. 3. 34 for Vy = 0. 

The maximum vertical rates of climb and minimum rates of descent in an /335 
autorotation regime are calculated in Table 3.11. 

Determination of the other flight data is accomplished by means of the 
graph in Fig.3.Z|B. 

A comparison of N^.f calculated by the auxiliary graph in Fig. 3. 2? with 
Mh. f found from helicopter performance and efficiency shows satisfactory agree- 
ment: Vv is also close in magnitude. 

''max 

Figure 3.49 shows a graph of required power based on Ga for helicopters of 
various configurations. The graph is calculated by means of the helicopter 

Nh t 
performance graph given in Fig. 3. 46. The ratio — -^ — is determined by the 

formulas : 

Qa 75iKi,r] K'^-'^U 

359 



in forward flight, and by 



in hovering flight, 



Ga 



1 

755 



/njMo 



TABIE 3.11 



(4.28) 



H 

V„, km/hf 
V« 
h 

fj, . 

^^ycL ,„. mlsic 




1430 
115 
0.163 
0.138 
1.010 
880 
550 
4.86 
—7.8 



1000 

1500 
115 
0.163 
0.152 
1.002 
865 
635 
5.57 
—7.6 



1860 
1550 

120 

0.170 

0.165 

0.989 

8ti7 

693 
6.0 

—7.4 



3500 


5000 


1315 


1380 


125 


125 


0.177 


0.177 


0.195 


0.229 


0.953 


0.850 


865 


970 


450 


410 


3.75 


3.05 


-7.2 


—7.2 



5500 

1300 
115 
0.163 
0.242 
0.850 

1060 
240 
1.83 
—7.9 



The power utilization factor § was taken as equal to O.93_for two-rotor 
helicopters, and as 5 = 0.88 for a single-rotor helicopter at V s 0.15 and as 
§ = 0.83 at V = 0. In hovering flight, the helicopter wing is swept by the 
rotor and creates a negative lift; therefore, at V = the lift coefficient of 
the rotor increases by 2$ for the single-rotor helicopter with a wing and by B% 
for the helicopter of side-by-side configuration with a wing; the value found 
from eq.(4.28) increases accordingly. 

We see from the graph that, because of a difference in 5 in hovering 
regime, the ratio N/G is lower for two-rotor helicopters than for single-rotor 
versions. 

The largest value of N/G refers to helicopters with a wing and with a re- 
duced rotor diameter, while the smallest value refers to helicopters with a wing 
and with reduced rpm for the single-rotor helicopter and for the side-by-side 
helicopter without a wing. 

Thus, to ensure the possibility of hovering, the helicopters in question 
should have a different engine power per kilogram of gross weight. Correspond- 
'ingly, they will have different flight data in forward flight. Table 3.12 gives 
some flight characteristics of helicopters which were obtained in our example /336 
at a flying altitude H = for the following characteristics of the engine: 



Nt. 



N, 



ho V » 



N„ 



= 0.85 Nt 



N= 



= 0.7 Nt. 



The maximum vertical speed of the helicopter was determined by the formula 



360 



V 



"dis 



^h,, 



y...= '^^^'^{-Oa Oa 



(4.29) 



The minimum rate of descent in an autorotation regime (Vy„),i„ was deter 
mined for \i, =0. 

TABLE 3.12 



Helicopter 
Configuration 



Side- by- aide 

Single- roior 

Fore- «nd- aft 

Si de- by- si de wi th 
wing and reduced 
di amc ter- 

Side- by- aide with 
wing and reduced 
peripheral speed 



Single- rotor with 
wing and reduced 
diameter 



Si n gl e- ro to r . wt th 
wing knd reduced 
peripheral speed 






U%-I 



tl4 



0.253 
0.284 
0.253 
0.306 

0.271 
0.315 
0.281 



301 
304 
280 
332 

321 
325 
317 



Ai 



A/ mlB 



K 



to 









-5 









0(47.7) 




0.456 
0.527 
0.694 
0.356 

0.348 
0.465 
0.44B 



6.6 

5.76 

2.61 

10.03 

9.02 
7.59 
7.06 



7.66 


246 


9.38 


244 


11.65 


185 


7.21 


280 


-6.24 


275 



— 9.17 



— 7.89 



268 



261 



0.127 
0.144 
0.169 
0.134 

0.121 
0.145 
0.133 





V Ga 


1 " 
Ihov 














Side- by- aide 


0.264 


306 





0.439 


7.19 


— 7.66 


252 


0.129 


Single- rotor 


0.264 


292 


29.4 


0.568 


4.67 


— 9.38 


230 


0.141 


Fore- and- aft 


0.264 


287 





0.667 


3.20 


—11.65 


203 


0.160 


Side- by- side with 
wing and reduced 
di ameter 


0.264 


310 


28.2 


0.413 


7.64 


— 7.21 


258 


0.126 


Side-by-side with 
wing and rediTced 
peripheral speed 


0.264 


317 


8.0 


0.357 


8.61 


— 6.24 


272 


0.119 


Single- rotor with 
wing and reduced 
diameter 


0.264 


300 


47.7 


0.655 


4.88 


-9.17 


235 


0.138 


Single- rotor with 
wing and reduced 
peripheral speed 


0.264 


309 


23.3 


0.477 


6.16 


-7.89 


251 


0.12f 



361 



The fuel supply required for flight over a given range at the cruising /337 
power of the engine was found by the formula 





G, _ 


IILC """" 


\,\LCe -77^ 
Ga 


n — 


I.ILC, ^ 


"z 


a 


Kr. 


^fi^fkra 



(4.30) 

The coefficient 1.1 is introduced to account for the fuel supply for navi- 
gation and fuel consimiption in transitional regimes for a height of H = 0. Here, 

Gf = -— — was calculated for L = 500 km at C^ = 0.32 kg/hp hr. 
G 

Table 3.12 indicates that, under equal conditions in hovering flight 
(Nt.o "= Nhov)» "the fore-and-aft helicopter has the worst flying qualities in 
forward flight: The rate of climb and cruising speed are appreciably smaller, 
the fuel requirement is greater, the rate of descent in an autorotation regime 

is greater, and -r— ^ — = 0.69, i.e., continuation of horizontal flight is pos- 
Nt. 

sible only if not more than one of the three engines fails. 



To improve the flight characteristics, the fore-and-aft helicopter should 
have a more powerful engine: N^.^ > Nhov* 

The side-by-side helicopter has the best flight data. 

The qualitative difference in flight data of helicopters of different con- 
figurations in forward flight also remains for helicopters of equal available 
horsepower per unit weight (see the second part of Table 3*12), and also if we 
take into account that helicopters of different configurations have a some^what 
differing flying weight at an identical lift capacity. At equal available power 
per unit weight for single-rotor and two-rotor helicopters, the former can hover 
only on a ground cushion, with its minimum speed outside the ground cushion 
being about 30 km/hr. 

By installing a wing on a single-rotor helicopter to reduce the rotor dia- 
meter by 11^ and increasing the engine power also by 11^ to ensure hovering out- 
side the ground effect, the flight characteristics of a helicopter in forward 
flight are improved: The maximum and cruising speeds increase by 20 km/hr and 
the- rate of climb by about 2 m/sec. Without an increase in engine power, the 
characteristics of a helicopter in hovering flight deteriorate, but in forward 
flight they change negligibly: The maximum and cruising speeds increase by 
5-8 km/hr and the relative fuel feed decreases by yfo. When a wing is in- 
stalled without changing the rotor diameter but with decreasing the peripheral 
speed, the flight characteristics of the helicopter improve both in hovering and 
in forward flight: The maximum and cruising speeds increase by 15 - 20 km/hr, 
and the relative fuel feed decreases by 8 - ^%. 

As noted above, the dynamic ceiling of a helicopter with a wing decreases. 
362 



h- Calculation of a Helicopter with a Tractor Propeller 



■ty Pt.p, i.e., the drag viill be equal to -^ + Qpar - Pt.p 



When calculating a helicopter with a tractor propeller (jet engine) de- 
veloping a thrust Pt.p, the drag of the helicopter must be reduced by the quanti- 

_G_ 
K 

The tractor propeller requires an amount of power determinable by the 
expression 

Therefore, the required power of a helicopter -with a tractor propeller /338 
is equal to 



or 



*•/ 270ET1 






K ' a a 



(4.33) 



In calculations using eqs.(4.32) or (4.33) we must assign the value of P^.p 
or Nt. p ; in so doing we must bear in mind that in steady horizontal flight the 
drag of the helicopter cannot be negative; consequently, the following condition 
should be satisfied: 

Accordingly, in a climbing or descending regime this condition takes the 
form 

^<^+-^^+^. (4.35) 

Such an additional term in formulas for Nh.f as in eq.(4.33) appears when- 
ever tractor propellers are installed on helicopters of any configuration. 

The flight data of a helicopter with a tractor propeller are determined in 
the same sequence as one without tractor propellers. 

Let us estimate how much the required power and maximum rate of climb may 
vary when a tractor propeller is installed on a helicopter. 

For Mo = 0.60 and average values of ty at high flying speeds (V ~ 0.35), 
the efficiency T] of the rotor can be considered equal to 0,87. Having taken 
? = 0.91 for a two-rotor helicopter and Tlt.p = 0.78, §t.p ^ 0.97 for the tractor 

propeller, we find the value in .parentheses in eq.(4.33): 1 - - * — — = 

= -0.05. 0.97 >^ 0.78 

363 



If the tractor propeller completely overcomes the helicopter drag, i.e., if 



J^-. 



K 



(4.36) 



then the increase in required power amounts to ^%, but if it overcomes this drag 
only by half, then the increase in required power amounts to 2.5%- Consequently, 
the losses are small. 

For Me = 0.7 and V = 0.35 - 0.4, when T] « 0.85 - 0.8, there can be a 
1-3^ gain in required power when a tractor propeller is used. 

In a maximum rate-of-clirab regime for Mo ==• 0.6 - 0.7, V = 0.2, and at 
average values of ty , the efficiency can be considered equal to at least T] = 
= 0.87. Consequently, if it is possible to obtain a very high value of tractor 
propeller efficiency (Tlt,p = 0.78) in a maximum rate-of-clin±i regime, then the 
value in parentheses in eq.(4.33) is equal to -0.05. For a side-by-side heli- 
copter, more than half of the available helicopter power is consumed in a maxi- 
miom rate-of-climb regime for producing vertical speed. Consequently, when all 
power is delivered to the tractor propeller, the total thrust of the helicopter 
decreases by 5^, whereas the excess of thrust used for climbing decreases by 
10^. The rate-of-clin±i loss will be about 1 m/sec. At T] = 0.9, we have 1]^, p = 
= 0.7, and when half of the available power is supplied to tractor propellers /339 
the rate-of-climb loss will amount to 2Q%, or about 2 m/sec. General considera- 
tions on when the installation of a tractor propeller or other propeller on a 

helicopter is expedient or necessary 
are given at the end of Section 7, 
Kt]1[ I I I J_J I I I I I I T^ I ] Chapter II. 



5. Comparison of Helicopter and Airplane 

At identical flying weight of heli- 
copter and airplane, the relation D = 
= Lw is approximately satisfied; this 
relationship is determined by design 
advisability of the size of wings and 
rotors. A comparison of the parameters 
of regular aircraft and helicopters 
shows that a wing has an aspect ratio 
larger by a factor of 7 - 9 and an 
equally larger load per square meter of 

area. Consequently, ( j = 

I'M aire 




Fig. 3. 50 Comparison of the Product 
of Aircraft Performance and Effi- 
ciency for Regular Aircraft and 
Helicopter. 






const. This means that a given aircraft and helicopter have 



an approximately equal induced drag at the same flying speed. 

Provided that D = Lh, the dimensionless coefficients of an airplane and 
helicopter are connected by the relation 



364 



G- 



'^i/. Y^'-^-^Vf-C*"^)'^. 



c,, = 



tyO JlXj, 



(4.37) 



For a comparison, let us take the following data of a helicopter and an air- 
plane. For a helicopter, a - 0.091; ty = 0.15, "c, ^ OtQfyi'i, For an airplane, 
the -wing characteristics are taken from data of wing exposirre to propeller slip- 
stream of a modern low-speed transport aircraft. A rectangular wing with a slat 
and double slotted flap was tested. The wing aspect ratio was X» = 9. The 
parasite drag coefficient c,^ ^ of the airplane, based on wing area, was taken 

to be equal to 0.025 [Sc^S of the airplane is approximately one half that of a 



helicopter: 



(Sc,S),ir, 



rot 



copter, Cx = 0.0075]. 



aire 



■'aire 
ro t 



0.025 

TT JU. 

4 



0.0036 and for the heli- 



TABLE 3.13 



7340 



Slat 



Slot 



Defl ection 
of Fl ap 



K„ 

" aire, 
^airc.'^t., 

i\.p =0-7) 

'^airc.'^t.p 
(tl^p=0.85) 



Retracted 



Overl apped 



Open 



20 40 50 



0.4 


0.8 


1.2 


1.45 


2 


2.4 


0.0185 


0.0439 


0.0873 


0.1175 


0.232 


0.417 


21.6 


18.25 


13.75 


12.35 


8.63 


5.77 


9.2 


11.6 


10.65 


10.2 


7.8 


5.44 


6.44 


8.12 


7.45 


7.15 


5.46 


3.8 


7.82 


9.86 


9.05 


8.67 


6.62 


4.62 



K.n^ 



0.479 


0.34 


0.277 


0.252 


0.214 


0.196 


— 


9.3 


8.2 


7.3 


5.9 


5.12 


— 


0.875 


0.915 


0.93 


0.965 


0.975 


— 


5.75 


6.03 


5.77 


5.1 


4.66. 


— 


4.43 


4.85 


4.72 


4.33 


4.00 



Moved 
Forward 



Open 



2.55 

0.451 

5.6 

5.36 

3.75 

4.55 



0.19 

4.19 

0.98 

4.5 

3.88 



SO 



2.87 

0.582 

4.93 

4.72 

3.31 

4.02 



0.179 

4.45 

1.0 

4.11 

3.62 



For the selected values of Cy^ we find, in Table 3.13, c^^^, K«, Y^%ira» 
Kairc'Ht.p. Foi" these same values of Cy , the values of V, K, T], Kh, and KuT] are 

365 



determined for the helicopter. Since the helicopter has additional engine power 
losses, the product Ki,T15, where % = 0.88, is calculated for a single-rotor heli- 
copter with a tail rotor. The graph in Fig. 3. 50 is constructed from the data of 
Table 3.13. 

A comparison of wing and rotor performance shows that, in the examined 
example, the wing without mechanization, at all Cy^^, has a performance greater 

by a factor of 2 - 1,7 than the rotor. The wing with mechanization has a 1^6% 
greater performance at Cy = 2.0 (at this c„, the minim.um flying speed corre- 
sponds to V„i„ = 0.214), whereas at (cy )aax "= 2.87 (Vnm = 0.18), the wing 
performance is only 10^ higher than that of the rotor. 

It follows from Table 3.13 and Fig. 3. 50 that the maximum value of ^e,iTe\.v 
^'^ ^t.p = 0.7 is by a factor of 1.75 greater than Kj,1i1§. The fuel consumption 
of the airplane per kilometer is less by the same factor than that of the heli- 
copter (at equal specific consumptions of the engine). The speeds corresponding 
to maxima of the product of the craft performance and efficiency is by a factor 
of 1.2 greater for the airplane (actually, airplanes fly the range at a greater 
speed and with a performance less than maximum) . 

On comparing a helicopter and an airplane at equal flying speeds, it will 
be found that at speeds reached by an airplane without the use of wing mechaniza- 
tion (V s 0.3 - 0.25), (K'n)air8 is by a factor of 1.5 - 2 greater than (K115)h. 
At V > 0.43, flow separation from the rotor blade begins at the helicopter. At 
low flying speeds, reached by an airplane with the use of powerful mechanization, 
(K'Tl)airo is less than (KTlg)^ owing to the large profile drag of the wing. Thus, 
it is aerodynamically less expedient to use such an airplane with its low at- 
tainable speed for long flights; a helicopter is then preferable. 

Table 3.13 shows that, at equal flying weight and at D = L^, the minimum /341 
speed of the airplane, determined by the quantity (cy^)nax = 2.87, will be V = 

= 0.18. A low flydng speed can correspond to this value of V only if the air- 
plane has a small wing loading. The minimum speed of the helicopter is deter- 
mined by the available engine power and is usually equal to zero, whereas_when 
the helicopter is overloaded it will not exceed a value corresponding to V = 
= 0.05. 

6. Power of Front and Tail Rotors in a H elicopter of 
Fore-and-Aft Configuration 

An expression was derived in Subsection 2 for determining helicopter per- 
formance and total required power of both rotors. However, the tail rotor usu- 
ally requires substantially greater power than the front rotor (by a factor of 
•1.5 and more). Let us derive an expression for determining the power required 
by each rotor separately. 

First, let us find the propulsive forces of the rotors. They are not 
identical, since various rotors may have different lifts and different angles of 
attack. 

366 



According to eq.(4«8), the relation between ax and 0^3 is equal to 

Oj = a, — ^Ont, -j- Ae„4 , 

where Ao^-oto is determined by eq.(4.14) with consideration of eq.(4«lS). 

Proceeding Itom the approximate expression for X:X = Y{a + a-x) , we find 

''1 

From the condition of equilibriimi of forces acting in the direction of 
motion (Fig. 3. 4), we obtain the following equality: 

x,+x^^r^^a„,,-]-Qfar=o. (4.39) 

In eq.(4.39), Qpar ^°^ ^ helicopter with a wing and tractor propeller re- 
presents the sum Qpar + ^w + 'i,, l^0!„ - Pt.p* 

From eqs.(4.3S) and (4.39), we obtain 

^1 = yr Qp^r- ^ AB^t, ; ( 4 . 40) 

1 + ^ 1 + -^ (4.41) 

Equations (4.40) and (4.41) indicate that vee-ing of the rotors by an angle 
Asrot redistributes the propulsive forces of the rotors, thus influencing the 

power required by the rotors. Owing to downwash of the tail rotor, its propul- 
sive force is greater (more negative) by an amount of YsAotots* ^'^ ^1 ^ Yg , the 
front and tail rotors do not furnish an equal share of parasite drag. 

Substituting Xi and Xg into the expression for calculating the required /342 
power 



j,j YiV / \ X^\ GV lYx X\ / , . „N 



we obtain 



367 



jg. 



It is easy to demonstrate that, on adding the expressions in parentheses in 
eqs,(4.43) and (4.44)> tihe sum will coincide with eq.(4-13) for the case of a 
fore-and-aft helicopter. The angle of vee-ing of the rotors Asrot ^.nd the 

Yi 
ratio at Tli = Tig will not influence the total power of the helicopter. At 

Yi = Ys and Aerot == Affrots* ^^^ power of both rotors is identical. It is 
necessary to note, however, that' the longitudinal stability of the helicopter 
deteriorates when Srotg > 0. 

7. Retraction of Landing Gear on Helicopters 

It is known that helicopters have a parasite drag about twice that of 
regular aircraft. 

This can be attributed to the specific configuration of a helicopter, the 
presence of a cabane and large hub of the rotor^ tail boom with a tail rotor 
placed high, and also to the necessity of loading and unloading in hovering 
flight and maintenance without the use of an airdrome. Therefore, various 
hoisting devices, numerous railings and hatches, movable doors, blisters, etc. 
are often installed on the outside of a helicopter. On the other hand, the 
weight coefficient which is lower than that of a regular aircraft necessitates 
a careful approach to any measures that reduce the parasite drag but increase 
the structural weight. 

Below, we will estimate the expediency of installing a retractable landing 
gear on a helicopter from the viewpoint of its load-carrying capacity. 

Retraction of the landing gear reduces the parasite drag of a helicopter by 

20 - 25%- Figure 3.51 shows the graph of the ratio — ^ — for two values of 



Cjt : 0.0075 and (by 23% less) 0.0056. The graph also contains the quantity 

— 7^ » 0.7 — f-i — -• It is obvious that, with retraction of the landing gear, 

the cruising speed of the helicopter increases from Vg^u = 253 km/hr (Mfj = 
= 0.207) to Vcru = 269 km/hr (Mn = 0.22), i.e., by 6%. The required fuel sup- 
ply will decrease by this same amount, while the range will remain unchanged. 
If the cruising speed is retained, then the required power diminishes by 9% 

368 



( -j=i — = 0.000555 in place of O.OOO605). Since the specific fuel consumption of 

turboprop engines greatly increases, upon a decrease in engine power, a change in 
power - as shown by calculations - will result in a change in fuel consimiption 
smaller by a factor of about 1,5. Consequently, we arrive at the same figure: 
The required fuel supply decreases by 6^. 

Now we can calculate that portion of the structural weight increase by /343 
retraction of the landing gear which is compensated by a decrease in fuel re- 
quirement. Thus, an increase in structu- 
ral weight by 1% of the takeoff weight 
will be compensated at a fuel requirement 
equal to 17^ of the takeoff weight, i.e., 
Gf = 0.17, since 6% of 17^ is 1%, A 
1,5^ increase in structural weight will be 
compensated ■when Gf =0,25. 



0.00075 



0.0005 



0.00025 




Fig. 3. 51 Ratio Nh . ^ /Ga of Heli- 
copter for Two Values of Para- 
site Drag Coefficient, as a 
Function of Mfi . 



The normal fuel supply of modern 
turboprop helicopters is about 15^ of the 
takeoff weight, with a maximimi of 20-25^. 
It is obvious that installation of a re- 
tractable landing gear on modern heli- 
copters is expedient if the increase in 
structiiral weight does not exceed 1 - 1.5^ 
of the takeoff weight. In so doing, how- 
ever, the maximum load-lifting capacity 
of the helicopter decreases in flights 
with a smaller fuel supply. 



A 6% decrease of fuel supply and an 
equal increase in cruising speed lead to cheaper hauling on helicopters, which 
should also be taken into account by the designer vjhen attacking the problem of 
landing-gear retraction. 

It should be pointed out that, on airplanes with a higher performance of 
the lifting system, a decrease in parasite drag will lead to a greater decrease 
in fuel consumption. Furthermore, regular aircraft have greater relative fuel 
supplies Gf , for which reason retraction of the landing gear on airplanes has 
become advantageous at cruising speeds lower than those presently used for heli- 
copters. 



Section 5. Aerodynamic Calculation of a Helicopter by the Power Method 

In an aerodynamic design of a helicopter by the power method, the condition 
of power balance in steady helicopter flight is used: The power supplied to the 
rotor is equal to the simi of all power losses. Thus, having determined all 
losses of power - both of the profile and induced type - produced in overcoming 
the parasite drag of the nonlifting parts as well as the helicopter weight com- 
ponent in climbing, we find the power which must be supplied to the rotor. 

The formulas for determining the torque coefficient of a rotor, derived in 

369 



Section 3, Chapter II, express the condition of the power balance. The same 
Section contains formulas and graphs for calculating the profile and induced 
power losses for a rotor. 

It is general practice to determine all power losses approximately so as /344 
to simplify the calculations; therefore, the aerodynamic calculations of a heli- 
copter by the power method constitutes an approximate method. 

1. Determination of Required Power in Horizontal 
Helicopter Flight 

The required power of a helicopter is equal to the sum of the profile and 
induced losses at the rotor and the loss due to overcoming the parasite drag of 
nonlifting parts of the helicopter 

^v =^^'- +^«^ +^^- (5.1) 

The profile power loss coefficient mpr is determined from special graphs, 
or by the approximate formula (3.75) from Chapter II: 

m,r=-\-'^.p,,il + 5V')P-\-Ant,,. (5.2) 

In dimensional form, the profile losses of a rotor are calculated by the 
formula 

^^^-^f-^p^^(^^r^^ (5.3) 

where § is the engine power utilization factor. 

The induced power loss coefficient is determined by means of eq.(3.83) 
from Chapter II 

m.^^^_^± =0.285-^, (5.4) 

vAiile, in dimensional form, the induced losses of a rotor are determined by 

The induced losses can also be represented as the product of the force of 
the induced drag of the rotor and the flying speed, or as the product of rotor 
lift and average downwash angle in the rotor plane and flying speed 

N. = '^ ^ind^^^ K(aa)K 
'"d 52 755 B2 755 ■ . (5.6; 

The average downwash angle in the rotor plane is determined by the average 
370 



induced velocity of the rotor 

2fb2qv 2fb2qv ' v5.7; 

Aa=— = — ^ . (5.8) 

V 2FB2QV2 ^^'°^ 

It is obvious that eq.(5.5) is also obtained from eqs.(5.6) and (5.8). 

It follows from eq.(5.5) that the induced power losses are directly pro- 
portional to the square of flying weight, referred to the effective linear di- 
mension BD (i.e., the span of the lifting system determines the volume of air 
flowing through the rotor). The induced losses are inversely proportional to 
flying speed and air density. 

Consequently, upon an increase in helicopter weight without a proportional 
increase in rotor diameter, the induced losses increase with respect to a /345 

n 

quadratic relation. If the load per unit rotor disk area p = -^ is retained, 
then the induced losses will be directly proportional to the flying weight and 
the ratio — -=; will remain unchanged. However, since increasing the helicopter 

tonnage causes p to increase (for decreasing the relative weight of the rotor), 

Nina 
the ratio — ^ is greater for heavy helicopters than for light ones. 



For multirotor helicopters, the induced power losses are determined as the 
svaa. of the product of the type of eq.(5.6), taken for all elements of the lift- 
ing system: 



^w=4"S"l'^^"- (5.9) 



In eq.(5.9), the downwash angles Aa are equal to the sum of all downwash 
angles for each element of the lifting system: the downwash angle due to self- 
induction defined by eq.(5.8) and the downwash angles due to interference*, 
whose expressions are given in Sections 1 and 4. 

As a typical example, let us develop eq.(5.9) for a fore-and-aft helicopter, 
using eqs.(4.14) and (4*18) for the downwash angle due to interference: 

^M=-^ ^-(l^iAai + KjAaj); 
.Aai = 



2/='B2qK2 ' 



■5S- For the terms containing the downwash angles due to interference, we can. take 
$/B^ = 1. 

371 



^°^^ OP-L\.o + 2\ 



2rB2QV2 ' <=» 2FB2QV2 • 
Having substituted Acci and Aofe into the expression for Nma, we obtain 







If the rotors have no excess, then the graph of k^o in Fig. 3. 8 furnishes 



Koo = 1 and 



^>n,=-:^,,^&^iy^+y2y- '"" °' 



m2B'iQVF ^ ' ' 75ilB4£ i/A£)2 

The expression for Nm^ shows that, at Koq = 1, the quantity Ni^^ does not 
depend on the distribution of helicopter weight between the front and tail 
rotors and is detennined only by the sum of lifts Yi + Tg = G. Displacement of 
the lifting elements along the direction of flight does not influence the quanti- 
"ty Ni„4, so that the expressions for N^a coincide for single-rotor and fore- 
and-aft helicopters. 

However, it must be borne in mind that, for a two-rotor helicopter of fore- 
and-aft configuration, the flying weight is equal to the thrust of the two rotors 
and that, at identical load on the rotor disk area p, the ratio G/D is twice 
that of a single-rotor helicopter. Therefore, as already indicated in Chapter I, 
Njni is by a factor of 4 greater for a fore-and-aft helicopter than for a single- 
rotor tjrpe, and the ratio — tt^— is twice as large. This explains why fore-and- 

G 

aft helicopters have poorer flying characteristics in horizontal flight and -vAiy 
the fl3ring characteristics deteriorate more noticeably upon an increase in /3h^ 

flying weight. 

If K,„ ^ 1, then N,„, = -i^i ^ ^t""" • 

For the side-by-side helicopter, the effective transverse dimension, i.e., 
the span of the system, increases with increasing flying weight, which is ex- 
pressed by the fact that Kg . s < 0. At Kj . g = (with the rotors spaced far 

apart), — p"'^- is the same for the side-by-side helicopter as for the single- 
rotor helicopter; at Kg. a < 0, the ratio — pr^— of the former is lower. 

G 

Losses for overcoming the parasite drag of the nonlifting parts of a heli- 
. copter are determined by the formula 

N^=S^=-^^V^^^^^- V^^ (5.10) 

'^ 765 16-755 12005 

or, in dimensionless form, 
372 



'"/«•= -<,^^V^, (5.11) 

where 

S = ~'-^^- ^5.12) 

When calculating a helicopter with a wing, we will refer the wing drag to 
the parasite drag of the helicopter, i.e., 

2^^ = (2-.5)^+('^.. +^..Aa,)5,, (5.13) 

where c,^ is the parasite drag coefficient of an isolated wing; Aa^ is the down- 
wash angle of the wing due to interference of the rotors. 

The interference of the wing with the rotors should be taken into account 
when determining the total rotor downwash angles for calculating Nm^ by 
eq.(5.9). 

Thus, the required engine power of a helicopter, in conformity with 
eqs.(5.l), (5.3), (5.5), and (5.10), is equal to 

N. =Np^ +-' 91 iSff^ ^3. ( 5 .14) 

If the helicopter has a propeller, then the power balance is expressed in 
the form 

The aerodynamic calculations can be performed in dimensionless form. In 
this case, the coefficient of required torque is determined, in conformity with 
eqs.(5.4) and (5.11), by the expression 

'"V ^ '"'"" + °'^^^ "^ " S "^ • ^5-^5) 

As a dimensionless form of calculation, convenient - for example - for 



the method of powers, is equal to 



comparative calculations, we can determine the ratio \* ' which, when using 

Ga 



Nl 



Ga~ 75£(<3,aMg) 'V 
where 

t^oMl= ^— 



'"f.x^'M?, 



(5.16) 


/347 


(5.17) 


373 



..3 '^Oy''M.lY 






(5.18) 
(5.19) 



At given p = -ri-, Cx = — s * height, and speed, the value of the products 

tyQMo and t^^^ jCrM© does not change upon a change in the rotor parameters uiR and- 

CT. Consequently, when studying the effect of rotor parameters on the magnitude 
of required power of the helicopter, eqs.(5.l6) and (5.18) are transformed into 

frif oM3= consta ■i-mpr<^M.l, 

As an example of the aerodynamic calculation by the power method, let us 
determine the required power of the Mi-4 helicopter. The helicopter data were 
given in Section 3. The calculation is performed by means of eq.(5.14) in 
Table 3.14. For simplicity, the profile losses were determined from the graphs 
in Figs. 2.63 - 2.66, using eq.(6.10) of Chapter II for converting t,; ; we can 
disregard the differences in blade profiles for Mo = 0.6. 



TABLE 3.14 
G = 7200 /t^. ; a = 0,063; « /? = 196 /w/sec! S = 0.84; 
/=• = 346 m 2; Mo = 0.6; ty = 0.138— ; N^r = 163-103 m^^A; 



N,-„^ = mO — ; Npar= 



2c^ 
1010 



V3A; 





//=1000/», 


<y =0.152; A = 


= 0,907 




V 


0.15 


0.20 


0.25 


0.30 


V, Mm/hr 


106 


141 


176 


212 


V, m/szc 


29.4 


39.2 


49 


58.8 


V 


4.5 


2.75 





-3 


Cx 


0.0088 


0.0089 


0.0092 


0.0096 


^CxS 


3.04 


3.08 


3.18 


3.32 


> 


—0.00314 


—0.00565 


—0.00915 


—0.0137 


0.0046 


—0.00135 


—0.0063 


—0.0117 


nipf 


•0.00286 


0.00306 


0.0034 


0.0037 


Npr 


420 


450 


500 


545 


Nind 


338 


253 


203 


168 


Npar 


70 


165 


340 


610 


V 


828 


868 


1043 


1323 



374 



2. Determination of Helic o pter Pe rformance Data /348 

The dependence of required power on flying speed is found by means of the 
formulas given in Subsection 1. Maximum and minimum speeds, maximtmi range and 
endurance, cruising and economic speeds are then determined by the method de- 
scribed in Section 3. 

During ascent, the propulsive force of the rotor increases by an amount 
equal to the projection of the helicopter weight onto the flight direction 
G sin Qfi.p. Consequently, the engine power of the helicopter increases by an 
amount of 

A^..c = 7^Gsin9,i,l/=^GK, (5.20) 

while the total power of the engine is equal to 

N^Npr^Nind^Np^+N^^. (5.21) 

Here, N^g o represents the variation in potential energy of the helicopter 
upon a change in its flight altitude. 

The components Np^, Ni„a, and N^^r o^i vertical ascent and in horizontal 
flight differ somewhat in magnitude. However, for approximate calculations we 
can disregard this, and under this assumption eq.(5.21) can be represented in 
the form 

It is obvious that the maximum vertical rate of ascent of the helicopter is 

y^_jm,-s_^^^t:l,,^^ (5.23) 



V 



max 



The discrepancy between eqs.( 5. 23) and (4. 24) can be explained by the as- 
sumption that Njnd 3-i^d Npr are equal in horizontal flight and vertical ascent. 
Equation (4»24) gives a more correct result. At flight altitudes where the 
rotor lift coefficient is less than the maximum value permissible with respect 
to flow separation, we can take, in conformity with the graphs in Figs. 2.166 
and 2.167, the average value of the propulsive efficiency as equal to 0.95 and 
determine V„ by the formula: 

ym ax *' 



"max 



0.95-75(Ar^ ,, -^*/n,m)S (5.24) 



At high flight altitudes in a climbing regime, where ty ~ ty^ , the speed 
Vy should be determined from eq.(4.24). 



B ax 



375 



3. Relation between Np,, Nj^ij, and Npar during Horizontal 
Flight of a Single-Rotor Hel icppt er 



It is of interest to examine the relation between individual components of 
the required power of a helicopter. Since the helicopter parameters determining 



N 



pr» ^ind* 



and N ar depend on the gross weight of the helicopter, we will give 



data for helicopters of different weight classes, 



Helicopters of different weight classes have a maximum weight coefficient 
at a different load per square meter of rotor disk area p and correspondingly 
have different peripheral speeds and solidity ratios, since the lift coefficient 
t limited in value by flow separation should be within 0.23 - 0.27 at the dy- 



namic ceiling and 0.13 - 0.17 near the ground. 



^I'P/Ms 



l^y - ■ — /, ight helicopter 

, 7~ — — — — AfediuB helicopte 



op t 
Mg=0.7 \ helicopter 




0.3 Mfl 



Fig. 3. 52 Quantity Nh.f/G and Relation 
between Components of Required Power of 
Helicopters of Different Weight Classes. 



376 



Let us assume that the charac- 
teristic parameters for a light 
helicopter with an all-up /349 

weight to G = 3000 kg are: 
Mo = 0.55, o- = 0.05; for a 
mediiom helicopter G = 
7000 - 14,000 kg; Mo = 0.6, o = 
= 0.07; for a heavy helicopter: 
Mo = 0.65 - 0.7, CT = 0.09. 

The quantity Ec^S of the 
helicopter referred to gross 
weight decreases upon an in- 
crease in weight owing to the 
relative decrease in the overall 
size of the helicopter (so- 
called "scale effect"). How- 
ever, with an increase in G of a 
helicopter the value of p will}, 
increase, while the parasite 
drag coefficient referred to the 



rotor area, Cj^ = 



G 



will 



change little for helicopters 
of different weight classes. 
Let us take it to be equal to 
0.0085 for light and medium 
helicopters and 0.0075 for 
heavy helicopters. 

For calculations at V ^ 0, 
we use eqs.(5.4), (5-ll), 
(5.12), and (5.16). The coeffi- 
cient mpr is found from the 
graph in Figs. 2. 63 - 2.70 as a 
function of the coefficients ty 
and tjt . At V = 0, we use 
eq.(8.27) and the graph in 
Fig. 2. 174. 



The calculation results are shown in Fig. 3. 52. Owing to an increase in p, 

a. Mo, the quantity — ^f-^ — in hovering flight is much greater for heavy heli- 

G 

copters than for light helicopters. 

In forward flight, eq.(5.l6) can be written in the form /350 

° '"' [ '^ -f P..AV "" ' ) (5.25) 

Nh f 
This expression indicates that, at high flying speeds, the value of — ^ , 

despite the increase in p and Mo, is lower for heavy helicopters than for light 

ECxS 
ones owing to a decrease in the ratio - — . 

If the available power of helicopters is equal to the required power in 
hovering flight, then the average value of the majclmum speeds of helicopters of 
different weight classes is equal to 210 km/hr (Mfi = 0.1?), 260 kmAir (Mf i = 
= 0.21), and 310 km/hr (Mn = 0.25). 

The graph shows that the profile power losses, in percentage of the power of 
horizontal flight, are 22 - 2?^ at V = and 50^ at average flying speeds, while 
they are k5% for light and medium helicopters and 55^ and more for heavy heli- 
copters at Vnaic* I't will be recalled that the graphs of mp^ in Figs. 2.63 - 2.70 
pertain to a rotor of average blade manufacturing quality and that Cxp of the 
profile increased by Ac^p = 0.002. 

The induced power losses amount to 73 - 7&% in hovering flight, 1+0% at 
average flying speeds, and only 13^ at maximum speed. 

Losses due to parasite drag amount to 15 - 10^ at average flying speeds and 
to 40 - 35^ at maximum speed. 

Thus, it turns out that, although helicopters of different weight classes 
differ in speed range, in load per square meter of rotor disk area, in peri- 
pheral speed, and in relative parasite drag, the power losses in fractions of 
the required power show a distribution that is practically the same at corre- 
sponding speeds. 

The above data permit an approximate estimate as to the degree of variation 
in required power of a helicopter on introduction of various modifications in 
the helicopter design. For example, an improvement in blade finish may cause its 
profile drag to decrease by 20^; consequently, the required power of the heli- 
copter will decrease by 10^ at medium and high speeds. In hovering flight, the 
required power diminishes by 5^, which is very substantial since, in this case, 
the relative efficiency of the rotor increases by a like quantity while the 

maximum thrust of the rotor increases by — k— ^ ^ = 3*3^ [the coefficient -75— is 

377 



obtained in accordance -with eq.(8.34) of Chapt.U]. 

Upon a change in blade shape, the induced losses of the rotor may vary 
within several percent. It is obvious that this substantially affects the maxi- 
raum thrust of the rotor in hovering flight but practically causes no change in 
the required power at high flying speeds. The change in blade shape at large Mo 
significantly changes the rotor profile losses (see Sect .3, Chapt.U). 

A 25% decrease in parasite drag of a helicopter leads to a 3% decrease in 
required power at medium flying speeds and to a 10^ decrease close to maximum 
speed; this yields an increase in maximum speed by 15 - 20 km/hr. 



378 



CHAPTER IV 7351 

ROTOR FLUTTER 

The phenomenon of rotor flutter has been a persistent companion of the de- 
velopment of helicopter construction. Numerous cases are known of the occur- 
rence of flutter in experimental helicopters during their first ground test or 
during flight tests. Cases of the appearance of flutter have been observed also 
during operation of helicopters that had already undergone all test stages. 
Rotor flutter has been the cause of a number of accidents. 

The greatest number of cases of flutter was observed at a time when this 
phenomenon had not yet been adequately studied and due attention had not yet 
been given to its investigation. At present, flutter has been studied in great 
detail, and there are numerous means for completely preventing its appearance. 
However, the helicopter designer must keep constant track of the rotor para- 
meters and hold them to limits that ensure the necessary safety margin before 
onset of flutter. However, these parameters vary constantly with design and 
technological modifications made in designing and plant testing of a helicopter 
and during its series production. Such variations continue even when the heli- 
copter has been placed in service. This is due to various circumstances. The 
most common case is deterioration of the individual blade balance, either due to 
penetration of moisture into the blade or due to its increase in weight during 
repair. 

Experience shows that even the slightest letting up in control of the rotor 
parameters will immediately cause appearance of flutter. This is primarily ex- 
plained by the fact that the designer strives to reduce the margin before onset 
of flutter to a minimum since the expenditures produced by an increase in rotor 
weight are generally proportional to the magnitude of this margin. Its increase 
requires a corresponding increase in blade weight or in weight of the structiiral 
elements of the rotor control system. 

As a result, the most economic design of a helicopter keeps these para- 
meters at the minimum level allowable by the flutter limits. Even their slightest 
variation produced by some unforeseen happening may lead to flutter. The blade 
parameters, at all times, are kept close to the flutter limit. 

This circumstance necessitated taking reliable measures to keep the flutter 
characteristics of a rotor within limits that would ensure prescribed margins /352 
before onset of flutter, which, as a rule, are rigorously standardized. These 
measures should be enforced both during production and service of the helicopter. 
In addition, each helicopter must be subjected to special ground tests to check 
for flutter in the final inspection. Experience gained in mass use of helicop- 
ters confirms the reliability of this inspection system. We can consider that, 
at present, conditions have been created that preclude the possibility of acci- 
dents owing to flutter. Actually, cases of unforeseen occurrence of flutter 
have almost completely stopped. 

379 



The achievement of the present favorable state as regards flutter was pre- 
ceded by extensive theoretical and experimental investigations. 

Valuable contributions to the development of the theory of flutter were 
made by P.M.Riz, L.N.Grodko, V.D.Il'ichev, M.S.Galkin, A.I.Pozhalostin, F.L. 
Zarzhevskaya, M .E .Lipskaya, V Jl.Pchelkin, and many other engineers. Numerous 
papers by foreign authors are also well known [see (Ref.39 - 42)]. 

Results of great importance for the development of the theory were obtained 
in flight tests on flutter carried out by S.B.Bren and A.A.Dokuchayev and per- 
formed by the pilot V.V.Vinitskiy. 

Many highly useful results were obtained by L.S.Popov, B.A.Kirshteyn, 
N.V.Lebedev, and B.BJ^artynov in tests of dynamically similar models. 

All this work led to rather clear and distinct concepts concerning the phe- 
nomenon of flutter which permitted developing new blades with the necessary 
parameter margins, without additional modifications after tests, as had often 
been necessary before. However, for this it was necessary, in designing the 
blade, to perform numerous rather laborious calculations. This Chapter will be 
devoted mainly to an account of the method of these calculations. 

In writing this C3iapter F.L. Zarzhevskaya was of considerable help to the 
author, for which the author extends his gratitute. 

Section 1. Basic Assumptions and Characteristics of an 
Approach to Flutter Calculation 

1. Bending and Torsional Vibrations of the Blade . 
Possible Cases of Stability Loss 

The theory of rotor flutter is developed on the basis of an investigation 
of bending and torsional vibrations of blades during their rotation in air. 

When solving the problem of bending and torsional vibrations of blades in 
air, the designer is interested primarily in two qualitatively different prob- 
lems. The first of them reduces to a determination of steady bending and tor- 
sional vibrations of the blade, occurring in all helicopter flight regimes. 
This problem requrres the development of special calculation methods which are 
a further development of calculation methods for forced vibrations of a blade 
and should, in particular, answer the problem of the effect of torsional de- 
formations of a blade on its bending vibrations and, accordingly, on the magni- 
tude of variable stresses from blade bending. The second question is associ- 
ated with a determination of the stability of blade motion. Usually, purely 
bending vibrations of blades are stable. Loss of their stability occurs only 
in flow-separation regimes. 

In studying bending and torsional vibrations we find that, at certain rotor 
parameters, there is a loss of stability of motion of blades which leads to 
flutter or divergence. The phenomenon in which blades undergo oscillatory in- 
stability is called flutter, whereas the phenomenon of aperiodic instability of 

380 



blade motion is called divergence. The most common of these two phenomena /^^'i 
in practice is rotor, flutter. Therefore, when examining bending and torsional 
vibrations of a blade, the designer is more interested in the conditions leading 
to flutter. 

2. Effect of Blade Attachment to Hub and the Possibility 
of TheoreticaT In vestigation of Flutter of an 
Isolated Blade 

The results of calculating flutter largely depend on the design configura- 
tion of the rotor and primarily on the conditions of blade attachment to the 
root, i.e., on hub design and rotor control system. The characteristics of 
blade attachment influence the boundary conditions of the problem and hence the 
design formulas for determining flutter parameters. 

The most common type of rotor with individual hinge ' attachment of each 
blade to the hub, with the control exercised over an automatic pitch control 
mechanism, will be examined below when presenting the method of calculating 
flutter. For rotors with a rigid and universal joint attachment of the blade to 
the hub or with some other type of control, the approach to flutter calculation 
remains the same. However, the conditions under which flutter occurs may change 
extensively. 

Flutter is greatly influenced by the design of the system controlling the 
angle of blade setting and, primarily, by the design of the automatic pitch 
control. The automatic pitch control couples the oscillations of the different 
rotor blades. Thus, as soon as this couple becomes sufficiently strong - and 
this generally takes place on real helicopters - it is impossible to investigate 
the flutter of an isolated blade. It is then necessary to study the flutter of 
the entire rotor a.s a whole. 

In all practical cases, there occurs only flutter of the entire rotor as a 
whole, when each advancing blade of the rotor duplicates the motion of the re- 
treating blade with some lag. Flutter of a single blade has never been noted. 

However, in many cases the investigation of flutter of a rotor as a whole 
can be reduced to calculation of the vibrations of an isolated blade. Therefore, 
calculation of the flutter of an isolated blade often furnishes a sufficiently 
comprehensive answer so that we can frequently restrict ourselves to this result 
in practice. In so doing, however, it is important to properly prescribe the 
stiffness of the isolated blade control. This question will be taken up in 
greater detail in Section 4» 

3. Different^. Types of Fl utter Differing with Respect to 
Blade Vibration. Flapping^ and^ending Flutter 

The problem of determining the conditions for occurrence of flutter is 
solved usually by means of differential equations of bending and torsional 
(binary) vibrations of the blade (see Sect. 6). These equations permit obtaining 
the parameters of different types of flutter which differ by the blade vibration 

381 



modes. The critical rpm and other parameters of flutter obtained from solving 
these equations are quite complex functions of the initial rotor parameters. 
Therefore, an analysis of these relations is conveniently begun with the simplest 
particular case. In fact, flutter in which blade vibration in the flapping plane 
occurs mainly with the fundamental vibration mode of the blade is most wide- 
spread in practice. Bending strains of the blade in this case have the charac- 
ter of an admixture to the vibration mode and do not determine the phenomenon. /354 
Therefore, in this case all relations of interest to the designer can be ob- 
tained from examination of a rotor model with blades that have absolute flexural 
rigidity and execute flapping vibrations about the flapping hinges. This type 
of flutter will henceforth be called "flapping flutter" in contrast to "bending 
flutter", whose characteristics cannot be determined without regard for the 
flexural deformations of the blade. 



4. Characteristics of the Torsional Vibration Modes of 
a Blade and Possible Correlated Assumptions 

The relation between torsional rigidity of the blade and the rotor control 
system in most modern helicopters is such that, in torsional vibrations, the 
blade turns mainly as a consequence of deformations of the controls (Fig. 4.1) » 

In this case, the setting angle cp 
of the blade element over its 
length, especially at the most 
effective portion from r = 0.5 to 
the blade tip, vary so insignifi- 
cantly that in flutter calcula- 
tions we can set, with a suffi- 
cient degree of accuracy. 



as 





— 








— 




























\ 


1 






~ 





















— 


■^ 


= 








' 






V 


-§t 


— 












— " 






\} 


r^ 




_, 


,^ 


■^ 









. — 


-■ 




r\ 




T 

w 




— 








:::= 






, 


^ 






















\ 






^ 








■ — 


















































































~~~ 
























- 






























- 


- 




































- 


— 


- 


- 












- 
















- 












































_ 























































1 





2 





3 


a 


* 


0. 


i 


0. 


f 


0. 


7 


0. 


8 


0. 


9 


1 









r 



q)=const. 



(1.1) 



The acceptance of this law 
of angle distribution cp is equi- 
valent to the assumption that the 
blade is absolutely rigid in 
torsion and executes torsional 
vibrations only as a consequence 
of deformations of the control. 
To have this assumption lead to the smallest possible error, we will introduce 
into the calculations the equivalent value of the hinge stiffness of the con- 
trols which takes into account the elasticity of the blade itself. 



Fig. 4.1 Typical Natural Vibration Modes 
of a Blade in Torsion (the Curves Refer 
to Three Different Helicopters) . 



Calculations made to substantiate this assumption show that it can be suc- 
cessfully used for all rotors for which the values of the angle a < 0.1+ - 0.5 
(see Fig.4«l)» which probably encompasses almost all existing helicopters. 

It should also be noted that the described character of the relation be- 
tween torsional rigidity of the blade and its attachment causes the axis about 
which the blade elements in torsional vibrations are turning to come close to 
the axis of the axial hinge. Hence, the position of the axis of blade stiffness 
in the examined cross section loses its significance. This circimistance permits 



382 



the approximate assiraiption that, in torsional vibrations, the blade elements 
turn about the axial hinge. 



5. Assumptions on Blade Oscillations in the Plane of Rotation 



7355 



There exists a definite coupling between blade vibrations in the flapping 
plane and in the plane of rotation. This coupling is due to two types of forces. 

The stronger is the coupling created by Coriolis 
forces. The weaker is the coupling due to aero- 
djmamic forces. 

Let us examine in some detail the forces 
coupling vibrations in the flapping plane and in 
the plane of rotation. 

During vibrations in the flapping plane, 
Coriolis forces arise which act in the plane of 
rotation 




(X =- - 2, 



'"yy m, 



(1.2) 



where 



Fig. 4. 2 Coriolis Forces 
Acting on a Vibrating 
Blade. 



y = rate of displacement of the blade ele- 
ments in the flapping plane (Fig. 4. 2); 

y'= angle of inclination of the blade axis 
upon deflection of the blade from the 
plane of rotation; 

m = mass of the blade element. 



During blade vibration in the plane of rota- 
tion, variable Coriolis forces are set up which act in a direction close to the 
direction of centrifugal forces. These forces stretch the blade and therefore 
should be taken into account in differential equations of blade vibrations, 
along with centrifugal forces. 

The Coriolis forces acting in the direction of the blade axis can be deter- 
mined by the formula 

N^=~2<iixtn, . . 

where x is the rate of displacement of the blade elements during vibrations of 
the blade in the plane of rotation (Fig. 4. 2). 

The Coriolis forces determined by eqs.(l,2) and (1.3) relate the blade vi- 
brations in the flapping plane and plane of rotation. 

The aerodynamic forces create an analogous coupling. 

If, in the flapping plane, variable aerodynamic forces associated with a 
change in the value Cy act on the blade, then the component of these forces 



383 



Q=OT (1.4) 

will cause blade vibration in the plane of rotation [the value of i entering 
eq.(l.4) determines the angle of inflow]. 

During blade vibration in the plane of rotation, the aerodynamic forces 
acting in the flapping plane will vary as a function of any variation in the 
relative velocity U. 

Thus, the presence of the described couples requires that blade vibrations 
in the plane of rotation be taken into account also in flutter calculations. 
However, calculations and experiments show that blade vibrations in the plane of 
rotation have an insignificant effect on the critical numbers of revolution of 
flutter. Therefore, in all calculations of flutter, blade vibration in the 
plane of rotation can be disregarded. We must also take into account that, in 
the absence of thrust at the blade, when the angle of inflow § is equal to zero 
(such a position is possible for an untwisted flat blade) and the blade is not 
deflected from the plane of rotation so that y' = 0, the terms of the coupling /356 
determined by eqs.(l.2) and (l.4) are absent. Thus, in this case there is no 
coupling between vibrations in the indicated plane. 

6. Determination of Aerodynamic Forces Acting on 
a Vibrating T^ofile 

The occurrence of diverging vibrations in flutter is caused by aerodynamic 
forces acting on the blade profile. Therefore, the basis on which these aerody- 
namic forces are determined is very important. 

In performing practical calculations of flutter, the method of determining 
aerodynamic forces based on the "steady-state hypothesis" is widely employed. 
In this hypothesis, it is assumed that, during vibrations of the profile, it is 
acted on by loads that are the same as those created if the flow pattern formed 
at a given instant of time were to be time-invariant. The use of the "steady- 
state hypothesis" for calculating rotor flutter yields quite satisfactory re- 
sults which are in good agreement with experiments. Therefore, our entire ac- 
count will be based on the results obtained under application of the "steady- 
state hypothesis". Refinements that can be made by taking unsteady flow into 
consideration will not be examined here. 

The use of the "steady-state hypothesis" leads to the following well-known 
formulas [see for example (Ref.29, 32, 33)] for determining aerodynamic loads 
acting on a vibrating profile of unit length: 



r-ic;,^u^[.-iy^{i-f)-^,];] 



w„„r=±ebu^l^-f^^_ci^^x 



'a.r- 2 



xf-i^+(f-f)f»]). 

384 



(1.5) 



■where 



T = aerodynamic force per unit length acting on the vibrating profile 
In a direction perpendicular to the relative flow velocity U; 
^e^er ~ torsional moment per unit length of aerodynamic forces acting rela- 
tive to the axis passing at a distance Xg from the profile leading 
edge; 
9 = angle of blade profile setting in the examined sections; 
y = rate of displacement of the blade elements in the flapping plane; 
xo = distance between profile leading edge and flexural axis, i.e., up 
to the point where the elements of the blade start twisting under 
application of a torque; 
Of = distance between profile focus or a.c. and flexural axis of the 
blade; in some formulas below [see eqs.(2.13) and (5.2)] we will 



also use the designation Of- 



R 



Equations (1.5) sor<3 obtained for a plane-parallel flow. Therefore, their 
use for determining the helicopter blade loading is approximate also in this 
sense, since the flow past the blade markedly differs from plane-parallel. 

It is convenient to make a slight transformation of eqs.(l.5) when using /357 
them for the helicopter blade, by Introducing certain additional simplifications 
and refinements. The relative velocity U of the flow past the profile can be 
approximately equated to its component U^ parallel to the plane of rotation of 
the rotor. It must also be considered that the other component of this velocity 
Uy directed perpendicular to the velocity U^ differs from y by the amount of the 
velocity of the air stream flowing through the rotor. Therefore, for a hell- 
copter blade, these formulas are generally 
used in the following form: 



<"/ 




— 
















/ 




0.3 















J 


/ 




- 




- 












0.2 






1 








-^ 








01 

























T=^\ clQb ]^^ul + U,U,^j--f)bU,<f^ ; 



m„ 



16 



Qb^U^^- 



o,T. 



(1.6) 



0.1 0,? 0.3 0.1 as as o.i as 0.9 1.0 m 

Fig. 4.3 Position of the Aero- 
dynamic Center on the Mach 
Number, for a NACA 230 Profile. 



The last term in the first equation in 
the systems (1.5) and (I.6) ordinarily has 
little Influence on the calculation results. 
Therefore, it can be neglected without re- 
sulting in substantial errors. 



In calculations of flutter under condi- 
tions of axial flow past the rotor in hover- 
ing flight or in operation of the rotor under ground conditions, the aerodynamic 
loads can be determined on the basis of the linear dependence of the aerodynamic 
coefficient on the angle of attack. This assumption is also included in 
eqs.(l.5) and (.16). However, under conditions of forward flight, especially in 
regimes close to stalling, this assumption becomes quite inaccurate. Therefore, 
a method permitting rejection of this assumption will be discussed below in Sec- 
tion 7. Refined formulas for calculating aerodynamic loads for this case will 
also be derived in the same Section. 



385 



As is knovm, the Mach n\imber M has a strong influence on the aerodynamic 
characteristics of a profile. To calculate flutter of a helicopter rotor it is 
especially important that M have a substantial effect on the position of the 
profile focus which, as will be shown below, greatly affects the critical revo- 
lutions of flutter. Therefore, in calculations for each blade radius, we must 
take the position of the aerodynamic center corresponding to the local value of 
M at this radius. Figure 4.3 gives the position of the a.c. as a function of the 
Mach number, for a NACA 230 profile. 

When calculating flutter in forward flight it must be taken into account 
that the local Mach mmiber varies relative to the rotor azimuth. This, in turn, 
leads to fluctuations of the position of the profile focus during each revolu- 
tion of the blade. In approximate calculations, this circumstance can be disre- 
garded . 

When using the calculation method presented in Section 7, fluctuations of 
the a.c. relative to azimuth can be accounted for without difficulty, which is 
one of the important advantages of this method. 



Section 2. Flapping . Flutt er_of_ari Is olated Blade with 
Axial Flow past tjhe Rotor 



7358 



1. Blade Model 



Flapping hinge axis 



Feathering hinge axis 



The parameters of flapping flutter can be determined with sufficient reli- 
ability from a calculation based on the following assumptions: 

l) The blade is absolutely rigid in bending and vibrates in the flapping 

plane like a solid body as a 
consequence of turning about the 
flapping hinge. 

2) The blade is absolutely 
rigid also in torsion and exe- 
cutes torsional vibrations, 
rotating like a solid body about 
the feathering hinge of the hub 
as a consequence of deformation 
of the control, presence of an 
automatic pitch control mecha- 
nism, and installation of a flap- 
ping compensator. 




Fig.4«4 Blade Model Used in the These assumptions lead to 

Calculation. the possibility of calculating 

a blade model with two degrees 
of freedom, determined by the 

variables 3 and cp (Fig. 4.4) • This model is usually called "semirigid". 



386 



2. Der ivat ion of Differential Equations of Flutter 

In this Subsection we will derive the differential equations of flutter for 
a model of an isolated blade. It will be shown below, in Section 4, that in many- 
cases the theoretical investigations of flutter of a rotor as a whole can be re- 
duced to an examination of the flutter of an isolated blade. Therefore, it is 
expedient to evaluate first the effect of various factors on the flutter of an 
isolated blade and to determine later (in Sect. 4) in what manner and in what 
cases these results can be extrapolated to a rotor as a whole. 

Let us construct the differential equations of torsional- flapping vibrations 
of an isolated blade. These equations can be derived by equating to zero the 
sum of the moment of all forces acting on the blade during its vibrations rela- 
tive to the flapping and feathering hinges of the hub. As usual, we will examine 
small vibrations for which all terms of the second order relative to small dis- 
placements of the blade can be neglected. 

To avoid needless complication of the equations, let us assume that the 
distance from the axis of rotation to the flapping hinge is equal to zero (tq = 
= O). Then, the condition of equilibrium of the moments of all forces relative 
to the flapping hinge can be written as 

R R R 

f m (rp - o^) rdr + o)2 J m (r? — ocp) rdr=^T rdr, (2.1) 



where /359 

B = angle of rotation of the blade relative to the flapping hinge; 
cp = angle of rotation of the blade relative to the feathering hinge; 
m = mass per unit length of the blade element; 
a = distance from the axis of the feathering hinge to the center of 

gravity of the blade element ; 
T = aerodynamic load per unit length determined by eq.(l.6). 

The integrals entering the left-hand side of eq.(2,l) can be expressed in 
terms of the moments of inertia of the blade relative to the horizontal hinge 
Ih.h and the centrifugal moment of inertia of the blade I^f : 





Icf =J mradr. 



\ 



(2.2) 



On introducing these designations into eq.(2.l) and referring all terms of 
this equation to Ih,h> 'the expression can be rewritten in the form 



p + (o2p--^(9- + co2cp) = -J-f rrrfr. (2.3) 

'h.h 'h.h J 

For the regime of axial flow past the rotor, the velocities entering 



387 



eq.(l.6) can be equated to 






=wr, 1 

=u)^X-r^, J (2.4) 



where X is the relative rate of flow through the rotor. 

Substituting eqs.(2.4) into eq.(l.6), and then eq.(l.6) into eq.(2.3), we 

obtain 

1 



where Yo is the mass characteristic of the rigid blade [see eq.(2.14)]. 

The values of the coefficients his, ^ig, d^^, and d^g will be given below 
[see eq.(2.14)]. 

The moment of external forces, relative to the feathering hinge, loading 
the system that controls the angle of blade setting, can be written as 

^cc. = -(9 + '«'<P) fa., +(fi + «>2p)4_^ -M^r-h 

«. ^, (2.6) 

-^ 

where 

la.i, = moment of inertia of the blade relative to the feathering or axial 
hinge ; 
In = moment of inertia of the blade per unit length relative to this 
axis; 
^aer ^ moment of aerodynamic forces per unit length relative to the axial 
hinge with this moment being determined by eqs.(l.6): 
Mfr = moment due to friction forces in the axial hinge of the hub. 

The moment acting on the control system, Meon ^^^ ^^ expressed in terms /360 
of rigidity or stiffness and deformation of the control system: 

Mc„„=Cc,„\, (2*7) 

where 

Y = angle of rotation of the blade relative to the feathering hinge due 
to deformations of the control system; 
Coon ^ stiffness of the control system. 

In order to express the value of y i^ terms of the setting angle of the 
blade sections, we put 

(p = 9-xp + Y, (2.8) 



388 



where 

9 = angle of setting of the blade sections prescribed by the control 

system; 
H = flapping compensator. 

The angle 9 is determined from the expression 

9 = 9o— SiSinij)— ejcosij), (2.9) 

where 

9(3 = angle of blade setting at the root for p = 0; 
9i and Q^ = angles of cyclic pitch control. 



It follows from eq.(2.8) that 



Y=o4-xfi — e. 



(2.10) 



Substituting y into eq.(2.7) and then eq.(2.7) into eq.(2.6) and referring 
all terms of eq.(2.6) to the moment of inertia of the blade relative to the 
axial hinge Ia.h> ^^ obtain 



Mt 



^+(/k,+'"')? + ^2.(P + '«=?) +^p' P+7^- 



'■'^ ' ^ch 



-it J^«'-'^^=i«+"^^ i J ^".^V^^^- 



(2.11) 



Here pt« is the frequency of natural vibrations in twist or torsion of an abso- 
lutely rigid blade in compliant control: 



Pt.- 



VI 



ypr 
a.tr 



(2.12) 



Substituting into eq.(2.1l) the value of ffiaor from eq.(l.6) and taking the 
resultant equation together with eq.(2.5), we obtain a system of differential 
equations of binary vibrations of a rigid blade: 




f -\r d22^<f + (y^^ + *22"'^) ? + ^2l'^ + ^2l'0f< + 



'0 J 



rdr. 



(2.13) 



389 



The coefficients entering eqs.(2.13) can be determined by the following /36l 
formulas : 



V,2 = — - — \mradr=ioC2i, 



Co] 



== i mrodr, 

Ia.h J 



2 y /,.;, J V 4 b) ' 

R 

d,A= ^ cj -f-[ br'^o^ dr, 

^ 'a.h •) 



"22 — 



X 



_1_ 

R 







*12 = 



R 

yT-\ brHr-\-ioC^^, 
■'A.A J 



1 



*22=i+Y^; 



'a.h J ^ 



Yo = 



c;Qbo.7R^ 



21 



(2.U) 



The coefficients of eq.(2.14) entering the differential equation completely 
determine the behavior of the blade in vibration. Certain comments are neces- 
sary relative to these coefficients. 

The damping factor d^^^^ of flapping vibration of the blade is determined 
only by aerodynamic forces since the moment of friction forces in the flapping 
hinge is relatively small. A quite substantial addition d,, due to friction in 
the feathering hinge enters the damping coefficient of the torsional vibrations 
of the blade dgg, in addition to aerodynamic damping. The effect of friction in 
the feathering hinge -will be discussed in greater detail in Section 3 of this 
Chapter . 

The coefficient d-^g entering the equation is small and not essential for 



390 



the final results of the calculation. Therefore, it can be disregarded in 
practical calculations. 



If the ratio of the moments of inertia io 



-a. h 



-, then we can /362 



Ih.h 1000 

also neglect the coefficient c^g. In so doing, the system of equations (2.13) 
is simplified even more. 

It is important to note that the effect of the position of the center of 
gravity of the blade element will appear in the calculation only upon a change 
in the coefficient 



1 mra dr. 







(2.15) 



3. Particular Solution_ of the^ Diff erential Equation 

It is not difficult to demonstrate that the expressions 



fj* = flo — a, cos <J» — 6j sinij;, 
9* = ?o — <Pi cos tl* —'^i sin i> 



(2.16) 



are a particular solution of the system of differential equations (2.13) and de- 
termine the undisturbed motion of the blade. If the swashplate of the automatic 
pitch control is set in a neutral position and if 9i = Qs = 0, then the particu- 
lar solution of these equations is constituted by the expressions 



9* = ?o- 



(2.17) 



4. Differential Equation o f Disturbed Motion 
Let us substitute into eq.(2.13) 

P = P* + Prf, 
9 = 9* + ?^, 



(2.18) 



where B^ and cp^ are the angles of deflection of the blade from a position cor- 
responding to its undisturbed motion. 

Then, bearing in mind that 3* and cp* represent the particular solution of 
eqs.(2.13), we obtain a system of differential equations of disturbed motion of 
the blade: 



9 + rf22">=<P + (/^^ +*22'"=) ? + C2,F+ d,,4^-(:'-Pl +^2l">=) P = 



Ac 



(2.19) 



391 



■•0 



In these equations, the subscript of the variables 3 and cp, designating 
that these vari^les refer only to disturbed motion, is dropped for simplicity. 

5. Notation of Differential Equations in Matrix Form 

It is convenient to vrrite differential equation (2.19) in the following 
matrix form: 

CX-\-D,>iX-\-{A-\-m'iB)X=Q. (2.20) 

Here, C is the inertia matrix: 

D is the damping matrix: /363 

l<., tin J 

A is the stiffness matrix; 

B is the matrix of centrifugal and aerodynamic forces; 

where bgi = Cg^ . 

X is the vector function: 

6. Solution of Differential Equations of Blade Vibrations 
Setting, in the system of equations (2.19), 






(2.21) 



we obtain the following characteristic equation: 

X< + A^l^+ (5,^24-52)'X2+(Ci«HC2)wX+D,»*4-D2;;^2=0. ( 2.22) 

392 



Here for simplifying the calculations, the values of X and cu are referred 
to the frequency of natural torsional vibrations of the blade pt«, i.e.. 



X = - 



''fu 



^/■w 



(2.23) 



The coefficients entering the characteristic equation (2.22) have the 
following form: 



A,. 






. "2~(''n~l~^22 — <^12''21 — ^2l"l2)i 

. ~(1 4" ^22 ~r ^11*^22 — ^I2<^21 — "12"21 — ^12^2l)l 



— (1-xr,.^), 



VJi 



. "2~ ('^22"r'^Il''22 — ''l2"21 — <'2l"l2)> 
0^21 



(cfii — xoTij), 



'0621 



— (^22~^12'^2l'' 

'0''2I 



(2.24) 



Let us examine the behavior of the roots of the characteristic equation /364 
(2.22) for different rotor parameters. 

In the major portion of the rpm range of practical interest, the motion of 
the blade is determined by two pairs of roots: 



^/ =gi±ipi'i/ic/'^// = q2~fiP2- 



(2.25) 



Figures 4.5 and 4.6 show the dependence of the real and imaginary parts of 
these roots on the rotor rpm and on the blade balancing. In both graphs, we 
plotted, on the abscissa, the rotor rpm n related to the frequency of natural 
vibrations of the blade in torsion pt^, expressed in oscillations per minute: 

The values of n coincide in magnitude with the values of the relative 
angular velocity 



393 



p,.pi 



to 



as 



as 



0.1 



0.2 



C„ ''0.8(?3.5''/o) 




"Iff (21%) / 



- - — - /■ 



C,,='3.0(*£S%) 



He I icop ter 



rpn range 



Jf 






A, 



m 



/4 



^ 



/y 



// 



N 



k 



/ 



/ 



r 



•\ 



^"\ 



Pi 



M. 



./ 



y 



-0(23%) 
C2,'=0.8{23.5%) 



\ 



N 



\ 



/- 


=!.6(i 


'I'M 


\ 






\ 








\ 






^ 






\ 






\ 





0.2 



01 



0.6 



Fig.4.5 Imaginary Part of the Roots 
of the Characteristic Equation as a 
Function of Angular Velocity, for 
Different Values of the Coefficient 
Cgi . [In this diagram, as well as 
in Fig.4«6, we indicate the absolute 
value of Cgi (without the minus 
sign).] 



where pt„ is expressed in rad/sec. 

Therefore, we will henceforth use 
the designations n and cju on an equal 
footing. 

The roots of the characteristic 
equation determine the law governing 
the motion of the blade after some 
extraneous action (in practice, this 
may be - for example - a gust of wind) 
unbalances the blade. In this case, 
the value of the real part of the root 
q determines the rate at -which the 
amplitude of the vibrations varies, 
whereas the imaginary part p determines 
their frequency. The negative real 
part of the root corresponds to damping 
oscillations of the blade. When this 
quantity is positive, vibrations of an 
amplitude increasing in time will be 
generated. 

The first pair Xj , shown in 
Figs. 4. 5 and 4*6 by broken curves, de- 
termines the motion in which deflection 
of the blade relative to the flapping 
hinge is ^predominant. The second pair 
of roots Aj J , shown by solid curves, 
determines the motion with an appreci- 
able rotation of the blade relative to 
the feathering hinge which is due to 
deformation of the controls. 



This second motion is of greatest 
interest since, at certain blade balancing, the real part of the root qg passes 
into the area of positive values (see Fig. 4. 6), which corresponds to vibrations 
of increasing amplitude, which are known as flutter. /365 

The values of the rotor rpm at which qg = are usually called "critical 
rpm of flutter". 

When qg < 0, the blade executes damping oscillations. In this case, the 
value of qs determines the magnitude of forces that produce damping of the blade 
vibrations and constitutes a criterion for their stability. It follows from 
Fig. 4. 6 that the damping forces begin to decrease long before the critical 
flutter rpm. This decrease is observed even when flutter cannot arise no matter 
what the rotor rpm but the margin for blade balance is insufficiently narrow. 
A decrease in aerodynamic damping, and hence of stability of blade vibrations, 
is undesirable and may have an adverse effect on the characteristics of heli- 
copter controllability. 



394 



-0.Z 




-0.B 



-f.O 



Fig. 4.6 Real Part of the Roots of the Characteristic 

Equation as a Function of Angular Velocity, for 

Different Values of the Coefficient csi . 



The peculiarities of the behavior of the first pair of roots \ will be 
examined below in Subsection 8. 



7. Determination of the Critical Flutter Rpm 

To determine the critical flutter rpm, it is possible to derive an ana- 
lytical expression if, in the characteristic equation, we set 



X==tyr,2=i^^^; (92 = 0) 



(2.26) 



Then, the characteristic equation (2.22) reduces to a biquadratic equation 



where 



2L = 



C, {2C2 — .4,62) + Ax (yli Os — B1C2) 



7366 

(2.27) 

(2.28) 

395 



k 



M-- 



Cj(C2-^,B2) 



from which we can determine the critical flutter rpm 

^flu=VL^Vl^^^^- (2.29) 

The vibration frequency in flutter is determined from the expression 



^,.-l/^^^- (2-30) 



8. Blade Divergence 



A study of the graphs in Fig.4.6- indicates the behavior of the first pair 
of roots Xj . 

Beginning with a certain rotor rpm, the imaginary part of this pair vanishes 
and two real roots appear. The presence of real roots indicates aperiodic mo- 
tion of the blade. 

With a further increase in rpm, one of these roots A.j ^^ passes into the 
region of positive values, which characterizes the appearance of aperiodic in- 
stability at this rpm, known as blade divergence. 

The value of the rotor rpm at which X. = is known as the "critical rpn 
of divergence" and can be determined by the formula 






Usually, the critical divergence rpn is higher than the critical flutter 
rpm and the maximum rotor rpn. However, in a number of special cases, blade 
divergence is a decisive factor. For example, the possibility of the occurrence 
of divergence does not permit using negative values for the flapping compensator. 
At H = 0, the possibility of occurrence of divergence is already quite real, and 
at small negative values of k the blade becomes aperiodically unstable. This 
circumstance must be taken into account when designing the rotor hub, especially 
when deflection of the blade relative to the drag hinge kinematically leads to 
a decrease in the values of k to below zero. 

9. Parameters Characterizing Blade Bala nce (Ef fective 
Blade Balance) " 

To evaluate a blade from the point of view of possible flutter, it is con- 
venient to introduce several concepts characterizing the position of the e.g. 
of blade elements over the blade length. The quantity 

396 



==f (2.32) 



is called blade balance in a given section. 

If balancing of the sections is constant over the blade length, then the /367 
value of the coefficient c^^ entering the equations will be directly related 
with the magnitude of this balance. The flutter characteristics of a blade in 
this case can be characterized by the value of the balance of its sections. 

In practice, however, balancing of blade sections lengthwise is always dif- 
ferent. Therefore, it is convenient to evaluate its flutter characteristics by 
means of the so-called effective balancing. 

The effective balancing of the blade in question is defined as the balanc- 
ing of some equivalent blade with an identical rotation of the centers of gravi- 
ty over the length and having the same value of the coefficient Cgi . It is con- 
venient to .assume the planform and mass distribution over the length of the 
equivalent blade as being identical to those of the blade in question. In this 
case, the effective balancing of the examined blade can be determined by the 
expression 



0»»r = 



R 

J mar dr 

— Cixla.h 



<//~" R — R 



f mbr dr J" mbr dr 



(2.33) 







For blades having the axis of the feathering hinge at a distance constant 

_ Xp 

in percent of the chord from the leading edge Xq = -^ — = const, it is convenient 

to characterize the effective balancing of the blade by the value of balancing 
of an equivalent blade relative to its leading edge 

-««//= -^0 +"«//• (2.34) 

Since the position of the axis of the feathering hinge has only a slight 
effect on the values of the critical flutter rpm, it is convenient to reckon 
effective balancing from the leading edge also in cases in which the condition 

-T- — = const is not satisfied. Then, the effective balancing can be determined 

by the expression 

R 

J mar dr 

^y/=y-— • (2.35) 

J mbr dr 



The effective balancing of manufactured blades can be determined only by 

397 



cutting the blade and experimentally determining the balancing of its individual 
segment s . 

10. Dependence of Critical Flutter Rpm on Blade Balancing 
and Values of the Flapping Compensator Coefficient 

To illustrate the effect of various parameters on the critical flutter rpn. 
Figs. 4.7, 4.8, and 4.9 give the results of calculations performed by eq.(2.29). 
The curves refer to different values of the flapping compensator coefficient k 
and to three values of the position of the feathering hinge axis xq at a constant 
position of the profile focus. 

The graph shows that a shift of the e.g. toward the leading edge, just as a 
decrease in the flapping compensator, will improve the flutter characteristics 
of the blade, whereas a shift of the e.g. toward the trailing edge and an in- 
crease in the flapping compensator will lead to a decrease in the critical 
flutter rpm. These results coincide qualitatively with experimental data. 





7368 



Fig.4.7 Critical Flutter and Divergence Fig. 4. 8 Critical Flutter and Diver- 
Rpm as a Function of Effective Blade gence Rpn as a Function of Effective 
Balancing, for Xq = 0.18. Blade Balancing, for Xo = 0.23. 

A comparison of the results of calculations performed for three different /369 
positions of the feathering hinge axis shows that the effect of this parameter 
on the critical flutter rpm is incomparably weaker than the effect of blade 
balancing. Consequently, the critical flutter rpn depends mainly on the mutual 



398 




za'/o 



2'^% 



25% 



26% 



^Vf 



Fig. 4. 9 Critical Flutter and Di- 
vergence Rpm as a Function of Ef- 
fective Blade Balancing, for 
xo = 0.28. 



position of the centers of gravity of 
the blade elements and of the profile 
focus. Therefore, a shift of the a.c. 
of the profile relative to the chord is 
just as effective as a shift of the 
blade balance. 



11. Blade_ Arrangement 

The presented dependences of the 
critical rpm on the balancing permit 
necessary" conclusions with respect to 
blade arrangement. It follows from the 
above calculations that the best way to 
improve the flutter characteristics of 
a blade is to shift its centers of 
gravity as much as possible toward the 
leading edge and to use aerodynamic pro- 
files which, in operating flight regimes, 
have their aerodynamic centers as far 
rearward as possible. This measure has 
a favorable effect even when the blade 
spar is shifted toward the leading edge 
to create forward balance, together with 
the feathering hinge axis which often is 
associated with the axis of the spar. 
The arrangement of the blade shown in 
Fig. 4 .10 is an example of such a solution. 



However, it must be borne in mind that the statement as to the relatively /370 
weak influence of the position of the feathering hinge axis on the flutter 
characteristics in comparison with blade balancing holds true only when the 

variation in these parameters is 
of the same order of magnitude. 
In practice, a shift in the posi- 
tion of the feathering hinge can 
be performed in appreciably wider 
limits than a shift in blade 
balancing. Therefore, this should 
be regarded as still another means 
of influencing the blade flutter 
characteristics. 

The blade whose arrangement 
is shown in Fig. 4.11 can serve as 
an example for the case in which 
a change of the position of the 
Fig. 4. 10 Blade Arrangement with Feather- feathering hinge is used as a 
ing Hinge Axis and Spar Shifted toward means of improving the flutter 
the Leading Edge. characteristics. 




399 



rm 



12. Effect of Control Rigidity 

A highly important parameter 
greatly influencing the flutter 
speed is the magnitude of the fre- 
quency of natural blade vibration 
in torsion or twist ptw . In the 
idealized blade scheme examined 
here, the magnitude of this fre- 
quency is completely determined by 
the hinge rigidity of the system 
controlling the angle of rotor 
setting Coon» I'^ practice, how- 
ever, the magnitude of this fre- 
quency is influenced also by tor- 
sional deformations of the blade 
itself. Therefore, to take into 
account the torsional rigidity of 
the blade it is proposed to use, in calculations by the approximate method pro- 
posed here, the value p^^ calculated with regard to deformation of both the con- 
trols and the blade. 




Fig. 4. 11 Arrangement of Blade with 
Turned Feathering Hinge Axis. 



We see from the differential 
the critical flutter rpm (flutter 



'«// 



26% 



P«% 



22% 



^one of possible 
flutter at low 
control rigidity 




equations of blade vibrations [eq.(2.19)] that 
speed) and frequency of vibrations in flutter 
are directly proportional to the 
quantity Pt„. Therefore, in all calcu- 
lations whose results are presented in 
the above graphs, the flutter speed is 
referred to p^^ and is characterized 
by the relative quantities 



Zone of impossibility 
of flutter 



OM 



0.8 



12 X 



Fig.4«12 Boundaries between Zones 
in which Flutter is Impossible and 
the Zone in which it Arises at 
Small Control Rigidity, 
o - Rotor blades for which no 

flutter was observed 
n - Rotor blades for which 
there was flutter. 



- n.iu 



(2.36) 



13. Conditions for Absence of Flutter 

The character of the dependence of 
flutter speed on various parameters 
shows that the creation of the neces- 
sary flutter characteristics does not 
require a simultaneous change of all 
parameters. Production of the necessary 
characteristics is possible upon satis- 
fying even one of the two following 
conditions: 



The first condition is the 
creation of a sufficiently high tor- 
sional rigidity of the blade and its 
attachment to the control system, so 
that 



/371 



400 



p. >Mn 



(2.37) 



Here, rinax is the maxlm.'um possible rotor rpm. It is sufficient that kx = 
= 4-5. 

When the condition (2.3?) is satisfied, there is no need to secvire any spe- 
cific transverse blade balancing. It can be arbitrary, and there is no need for 
introduction of special counterweights into the design. 

The second condition is the creation of a sufficiently forward blade balanc- 
ing so that 



V/<-^" 



h'lf 



(2.38) 



Here, Xna is some limiting blade balancing at which flutter is impossible 
no matter how small, say, the torsional rigidity of the blade attachment to the 
control. 

Figure 4«12 gives the calculated value of the limit balancing Xii„- as a 
function of the value of the flapping compensator and position of the feathering 
hinge axis 5Ea. This balancing divides the entire area of parameters into two 
zones, in one of which flutter cannot occur even at very low control rigidity. 

The available statistics on full-scale flutter tests of rotors with blades 
whose effective balancing after the test was determined by cutting, satisfacto- 
rily agree with the described limit ( see 
Fig. 4. 12). 

Experience has shown that, in 
practice, only one of the indicated con- 
ditions is more readily met. 



14. Mechanism of Generation of Forces 
Exciting Flutter 




Fig. 4 .13 Damping Forces Acting on 
a Vibrating Profile without Tor- 
sional Vibrations of the Blade. 



ing of this phenomenon. 



The calculation methods that reduce 
to a determination of flutter parameters 
are^ left without an explanation of the 
mechanism of action of aerodynamic forces 
leading to the generation of divergent 
vibrations. A detailed study of the 
nature of the aerodynamic forces acting 
in flutter yields no new data for flutter 
investigations. However, in certain 
cases it does promote better understand- 



Let us examine the blade model which was described in Subsection 1 of this 
Section. For simplification of the problem, we will limit ourselves to the 



401 



particular case vjhere the aerodynamic center coincides vrith the axis of the 
feathering hinge and -hAiere Of = 0. We can also disregard the dependence of the 
force T on cp, which does not have any particular meaning. Then, the aerody- /372 
namic forces acting on the profile can be represented in the form 

T = ±-c'^QbU^a; (2.39) 



2 y 
Te 

where a is the angle of attack of the blade element. 



^..r=-^Qb'Ui (2.40) 



The moment of the aerodynamic forces T acting relative to the flapping 
hinge can be written as 

^*.A=*°' (2.41) 

where 

/? 



'- = ~c;Q<.^jbr^dr. (2.42) 



We will assiime that the blade executes vibrations relative to the flapping 
hinge according to the law 



? = PoSinp(. (2.43) 



In this notation, the time reference point is taken from the instant at 
which P = 0. 

First, we will examine the case in which the blade does not execute tor- 
sional vibrations. The angle of setting of its elements will be considered as 
equal to zero and constant in time. In this case, the angle of attack of the 
blade elements will vary according to the law (Fig. 4.13) 

a=acos pi, (2-44) 

where 

"=-^»f- (2.45) 

The moment of the aerodynamic forces relative to the flapping hinge will 
vary by the same law 

M^j,=M cos pi. (2.46) 

In accordance with eq.(2.4l), the sign of M will coincide with the sign 
of a. 

If a < 0, as occurs in the case in question, then the moment relative to 
the flapping hinge always acts opposite to the angular velocity of blade vibra- 
tions P (see Fig. 4. 13) and does negative work in blade displacements. 



402 



The magnitude of this work diiring the vibration period can be calculated 
by the formula 



2it 

p_ 



A = ^M^I^^dt=.^M%pQ.o%'^ptdt=n%M, (2.47) 



Ott 

where T = ■ is the period of blade vibration. 

P 

The sign of the work A coincides with the sign of M which, in turn, coin- /373 
cides with the sign of a. In the examined case, A < 0. 

This means that the air stream flowing past the blade absorbs the work ex- 
pended to maintain blade vibrations. Thus, in the presence of aerodynamic forces 
the blade will vibrate with a constant amplitude Pq only if energy equal to the 
magnitude of work calculated by eq.(2.47) is furnished to it from without. 
Otherwise the kinetic energy of the blade and, together with it, the amplitude 
of oscillations Po > will diminish and the oscillations will decay. 

A different picture may be produced in the presence of torsional blade vi- 
brations. Torsional vibrations of the blade arise as a consequence of deforma- 
tions of the control system and kinematic coupling across the flapping compen- 
sator. Deformations of the control system arise from aerodynamic and inertia 
forces acting on the blade during its flapping vibrations. 

Centrifugal and inertia forces arising during .flapping vibrations of the 
blade create a moment relative to the feathering hinge due to the presence of 
an arm between the centers of gravity of the blade element and this axis 



'nmtrt= ~{p^ — w'^)?o^ mar dr Sin pt. (2.48) 



The aerodynamic forces create a moment on the arm between the profile focus 
and the feathering hinge axis Of 

R 

fna<ir= clQuip'pA br^oj, drcospi. (2.49) 



At Of = 0, this moment is equal to zero. Therefore, as a consequence of 
flapping vibrations only the moment mj„ej.t will act on the blade. Under the 
effect of this moment, the blade pitch control is deformed and the blade begins 
to execute torsional vibrations. However, the phase of the torsional vibrations 
will not coincide with the phase of the flapping vibrations. Phase shift of the 
torsional vibrations is caused by damping forces acting in the control system 
directed opposite to the vibrations. These forces are caused by forces of aero- 
dynamic damping determined by eq.(2.40) and by the moment of friction acting in 
the feathering hinge of the blade. The direction of phase shift of the torsional 
vibrations depends on the sign of the external moment mj^gyt* 

The law according to which the blade executes torsional vibrations 

403 



(Fig. 4. 14) can be written as 



tp = 9 cos pt -\-os'm pt. 



(2.50) 



Here, it is assumed that the initial setting of the blade elements is equal 
to zero. 

The angle of attack in this case will vary according to the law 

a =a cos/?/ -fa sin /7^, (2.51) 



where 



«=?-8o— ; 



a-=f. 



The appearance of a sinusoidal component in the law of change of the angle 
of attack a and, along with this, the sinusoidal component of the moment relative 
to the flapping hinge, does not influence the energy transfer during blade /374 

vibrations. Actually, a check on the 
work done by the sinusoidal component 
of the moment H in blade displacements 
relative to the horizontal hinge will 
show that it is equal to zero: 




A^^^Msm pt p^Q cos pt dt - 





=0. 



(2.52) 



Fig. 4.14 Damping Forces Acting on 
a Vibrating Profile in the Presence 
of Torsional Vibrations of the 
Blade . 



The magnitude of the co sinusoidal 
component of the angle of attack a, as 
follows from eq.(2.5l), largely depends 
on the sign and magnitude of cp. 

When 9 < 0, the work absorbed by 
the air stream flowing past the blade 
increases which causes a rise in the 
rate of damping of the free vibrations 
of the blade. Thus, when cp < 0, the 
stability of flapping vibrations of the 
blade increases. When 9 > 0, the work 
absorbed by the stream past the blade 
decreases and vihen 



^=^0^- (2.53) 

it becomes equal to zero, whereas when 

^>^°f- (2.54) 

the cosinusoidal component of the horizontal hinge moment is directed along the 
404 



angular velocity of the flapping vibrations 3 . This leads to "resonant build-up" 
of the blade. The kinetic energy of blade vibrations begins to increase, which 
leads to a rise in the vibration amplitude. Such a type of oscillation at 
amplitude build-up is known as flutter. 

Thus, the occurrence of flutter is associated with the magnitude and sign 
of the component of torsional vibrations cp. 

Let us examine how the quantity cp changes under the effect of an external 
moment varying by the sige law in conformity with eq.(2.48). Figure 4.15 shows 
the dependence of cp and cp on the vibration frequency p of the external moment 
miner t. As usual during vibrations close to resonance, the component 9 which is 
in 90° phase with the external forces first increases, whereas the vibration 
component coinciding in phase with the external forces changes its sign in 
resonance, passing through zero. 

Thus, the value of cp increases especially upon approaching resonance with 
the frequency of natural blade vibration in torsion. Therefore, flutter always 
occurs with a frequency close to but slightly below the frequency of torsion. /375 
Usually the frequency of flutter amounts to about 0.8 Ptw* 

It follows from the foregoing that flutter occurs as a consequence of the 

following causes: The torsional moment 
due to inertia forces acting during 
V',f° 7 11 I ( I l/TM I I \ \ 1 — I flapping vibrations of the blade leads 

to the appearance of torsional blade 
vibrations. In so doing, the torsional 
vibrations with a 90° phase shift rela- 
tive to the flapping vibrations increase 
especially strongly at frequencies 
close to the frequency of the natural 
vibrations of the blade in torsion. 
This component of the torsional vibra- 
tions leads to excitation of flapping 
vibrations of the blade. As soon as 
this excitation [first term in 
eq.(2.55)] becomes stronger than the 
forces damping the flapping vibrations 
[second term in eq.(2.55)], flutter 
will occur. 



-2 
-3 
-U 









- 














'S 


i— 


^ 


i~ 


r^ 
















V 


























\ 
















- 












\ 












9- 












i 


























^ 










\ 










- 


- 






- 


















) 


r-- 


— 


~~ 




















/ 


\ 




























/ 




\ 




















/ 








/ 






\, 






/ 














\ 


S 


~^ 


/ 


/ 






1 










s 








i 
1 




























^ 




y 










I 
















1 


















"^ 


zoo 








100 








eoo 




P 


s. 


> 


\ 








1 

/I 


































; 

1 
















—, 


p 


















\ 


k 




t 


































.y 























































Fig. 4. 15 Variation in the Torsional 
Vibration Components cp and cp during 
Blade Vibration Frequency. 



From the expression for the co- 
sinusoidal component of the angle of 
attack 

5=»-Po-^ (2.55) 

it is also possible to trace the effect 
of rotor rpm on flutter. Actually, the 
second term in this formula rapidly 
decreases with increasing rotor rpm, 
whereas cp does not greatly depend on 



405 



the rpE since the external torsional moment xa^j^ert i^ determined mainly by the 
vibration frequency [see eq.(2.48)] because of the fact that, during flutter, p^ 
usually is by a factor of 5 - 8 greater than cu^. The variation in 9 -with respect 
to rotor rpm is related mainly -with an increase in aerodynamic damping at increas- 
ing ou. 

Thus, on tracing the mode of variation of the quantities entering eq.(2.55) 
with the rotor rpm, it will be found that, at some value of w, the cosinusoidal 
component of the angle of attack a changes in sign and becomes positive. This /376 
leads to the appearance of flutter, beginning with some specified rotor rpm. A 
rearward shift of blade balancing leads to an increase in the absolute value of 



m. 



Inert 



[eq.(2.49)] and hence to an increase on 9. In this case, as follows from 



eq.(2.55), flutter arises at smaller 



u>. 




In the same manner, it is possible to trace the effect of various other 
parameters on the flutter speed. However, there is no need for this since this 
has already been done above with sufficient detail. 

Section 3. Consideration of Friction Forces during Flutter 

1. Character of the^ffect of Friction_ j^orce£ du ring Flutter 

The occurrence of flutter leads to the appearance of oscillatory motions in 
the hinges of the rotor hub and in the hinge control. Therefore, the friction 
forces acting in these hinges have a substantial effect on the critical rpm and 

on the nature of generation of flutter. 
Of primary importance in this case is 
friction in the feathering hinge of the 
blade loaded by a centrifugal force, in 
comparison with which the friction in 
all other hinges can be neglected. 

Experiments show that forces act- 
Fig. 4. I6 Recording of the Moment ing in the feathering hinge are similar 
of Friction in the Feathering in character to forces of dry Coulomb 

Hinge during Torsional Blade friction [eq.(4.l6)]. The introduction 

Vibrations. of these forces into the calculation 

makes the problem of flutter essential- 
ly nonlinear. Therefore, in simplified 
calculations it is natural to use any of the possible methods of linearization 
of friction forces. A more exact solution to this problem without such lineari- 
zation will be given in Section 7 of this Chapter. 

As is known, linearization of friction forces leads to the dependence of 
the damping coefficient on the amplitude of oscillations. Here the nature of 
flutter generation, described on the basis of the calculation changes at in- 
creasing amplitude, approaching that observed in experiments on helicopters. 
These results permit explaining numerous peculiarities in the development of 
flutter in full-scale experiments. The possibility of interpreting these char- 
acteristics appreciably facilitates the conduction of tests. 

406 



2. Lin earization of Friction Forces 

Let us use the energy method of linearization of friction forces. For this, 
we will replace the moment of friction acting in the feathering hinge of the 
blade by some equivalent moment whose magnitude is proportional to the rate of 
angular blade displacement 

^f.^=-Y,.9. (3.1) 

The value of the coefficient Yfr is determined from the condition of equali- 
ty of the work done during the vibration period by the moment of friction, 

A^r=4M^rVfio (3.2) 

and by an equivalent moment whose magnitude is proportional to the vibration /377 
rate 



^^?="V^//^/"' (3.3) 



where 

Mfr = constant (in magnitude) moment of friction acting in the feathering 
hinge, always opposite to the 'rate of relative displacement; 
cpfiu = amplitude of torsional blade vibrations in the feathering hinge 

during flutter; 
Pfiu = frequency of blade vibrations during flutter. 

The moment of friction acting in the feathering hinge can be considered 
proportional to cu^, since its magnitude is determined mainly by the centrifugal 
force 

In a number of cases, however, this dependence is disturbed as a conse- 
quence of the following circumstances: 

1) The bearing is Installed with appreciable prestressing. In this case, 
the load acting on the bearing is determined not only by centrifugal force but 
also by the Initial tension. 

2) The design of the packing glands is such that they have an appreciable 
moment of friction regardless of the magnitude of the effective centrifugal force. 

3) The use of too heavy a lubricant in the bearing creates an appreciable 
additional moment due to the generation of viscous friction. The appearance of 
relatively large viscous friction forces is often observed at low negative tem- 
peratures of the ambient air. 

All these facts have an influence on the flutter speed but introduce no 
fundamental features into the pattern of the phenomenon. Therefore, in the 
following account we will take eq.(3.4) as the basis. 

The coefficient af^ entering eq.(3.4) is determined from the expression 

407 






<^f=K^a^^ (3.5) 

where 

Sa.r - static moment of the blade relative to the axis of rotation; 
r^g = radius of the thrust bearing; 

f = coefficient of friction in the bearing. 

The values of the friction coefficients f are usually quite stable and 
amount to about 0.003 for ball and 0.006 for roller bearings. 

After equating eqs.(3«2) and (3.3)» we obtain the expression for determining 
the coefficient Yf r • 

^f'^^ "^lii^ ' (3.6) 

With this method of linearization, consideration of the friction forces 
leads to only one change in the initial equations (2.19), namely of the coeffi- 
cient dss standing for the first derivative of the angle of rotation of the 
blade in the hinge, which is supplemented by some addition d, ^ . 

In an investigation of flapping flutter with a blade rigid in torsion, this 
supplement should be determined by the formula 

^F = -r^ • (3.7) 

3 . Determination of Flutter Speed with Considerati on of F riction /378 

Equation (3.7) derived above, which determines the magnitude of the addi- 
tion term due to friction forces to one of the coefficients of the equations of 
blade vibration d^s, is distinguished by a highly important characteristic. 
This addition depends on the amplitude of blade vibration in the feathering 
hinge during flutter cpfiu* Consequently, the critical rpm at which the ampli- 
tude of oscillations theoretically remains constant in time depend on the ampli- 
tude of flutter oscillations. 

Figure 4.17 shows such a dependence for three values of blade balancing ob- 
tained in a calculation of flapping flutter. Along the abscissa in this diagram 
is laid out the amplitude of angular blade vibrations in the feathering hinge 
cpfiu, and along the ordinate the critical flutter rpa referred to the frequency 
of natural vibrations of the blade in torsion Hj lu . 

These curves determine the amplitude of the oscillatory regime, which forms 
the boundary between oscillations with amplitude build-up and damping oscilla- 
tions. 

For all practical purposes, this means that for flutter to occur some 
initial impetus is needed leading to deflection of the blade from a position of 
equilibrium by an angle determined by these ciicves, usually called the excita- 

408 




Fig. 4. 17 Critical Flutter Rpm as a 
Function of the Vibration Amplitude 
9f lu • 



quencies. 



tion threshold. 

If there is no such impetus present, 
flutter will not occur at all no matter 
what the rotor rpm' might be. 

For a comparison. Fig. 4. 17 shows 
the critical rpn for the case in which 
the moment of friction in the feather- 
ing hinge is Mfj. = 0. 



4. Effect of Forced Motion in the 
Feathering Hinge 

Quite a different picture of the 
occurrence of flutter is observed when 
forced motion is present in the feather- 
ing hinge of the hub caused by tilting 
of the swashplate of the automatic 
pitch control or by forced flapping 
vibrations of the blades arising in 
flight during oblique flow past the 
rotor. In this case, the vibrations 
in the feathering hinge following the 
occurrence of flutter are generated by 
a complex law consisting of two oscil- 
latory motions with different fre- 



Figure 4«18 shows, as an example, the pattern of this motion observed 
during flutter under conditions of ground tests when forced motion is present in 
the feathering hinge caused by tilting of the swashplate of the automatic pitch 
control (curve cp^ or ) ^^'^ 'the motion caused by flutter (cpfiu). /379 







■/?V^^ 



Oscillation period 




Fig. 4. 18 Character of Flutter in the Presence of 
Forced Motion in the Feathering Hinge. 

For convenience of further discussion, we plotted the rate of vibration in 
the feathering hinge rather than the displacements. 

The work of the friction forces acting in the feathering hinge can be 



409 



determined by the expression 

t 
A 



^r=]Mf,kdt, (3^gj 



where the moment of friction Mf^ is always directed opposite to the rate of 
angular displacement of the blade 9. 

If the rate of angular motion cp is the sum of two oscillatory motions 

then the work of the friction forces can always be represented as consisting of 
two works, in each of these motions 

^/' = V+^/''" (3.10) 

where ^ 


t 
^flu=^^f9fl„dt. 

ti 

Here, the moment of friction - as usual - is directed opposite to the rate 
of total motion cp. 

The simultaneous presence in the feathering hinge of two oscillatory mo- 
tions of different frequency always leads to the appearance of time segments 
during which the friction force coincides in direction with the rate of one of 
these motions, in this case doing positive work. In Fig. 4.18 the area segments 
corresponding to the positive work of friction forces in displacement of one of 
the composite motions of a frequency Pf lu are hatched. As a result, the overall 
magnitude of work of the friction forces during the vibration period in dis- 
placement of each of the composite motions decreases in comparison with the case 
where there is no concomitant motion. As applied to otir case, this means that 
the work expended for damping flutter vibrations markedly drops because an ap- 
preciable portion of the friction forces is expended by forced motion. This 
drop can be characterized by a special coefficient which represents the ratio /380 

J Aff 

where 

Af lu ~ work of friction forces during the vibration period in displace- 
ments of the component of motion caused by flutter, which is of 
interest here; 
Afr = work of friction forces during the same period when there is no 
concomitant forced motion. 

Figure 4.19 shows the dependence of the coefficient Afj^ on the amplitude 
410 



ratio of the velocity components of oscillatory motion ( ~) * 

^for _ 

At the values of ^^'■" = 1.5 - 2.5 of interest to us, the coefficient Afi„ 

Pf or 

depends little on the ratio of these frequencies. 

If the value of the coefficient Yfr is determined in this case, as was done 
above, from the condition of equality of work [see eq.(3.6)], then eq.(3»7) takes 
the following form: 



4a/r-^//a (0 1 



«4.A. Pfiu yiai 



(3.12) 



the 



It follows from this expression that the critical flutter rpm depending on 
coefficient d,, is related with the amplitude of forced motion in the axial 



fiflu 























— 






-'- 






— 


7 


^ 




















1 




















/ 




















/ 


/ 




















/ 














ff^ 






/ 




















/ 




















/ 





















OM 



0.8 



1.2 



I.S 






Fig. 4 .19 Dependence of the 
Coefficient Sfj^ on the Ampli- 
tude Ratio of the Velocity 
Components of Oscillatory 
Motion. 



hinge cp. 



*-nu 



since d,, depends on the quantity 
With consideration of the nonlinear 



dependence Af ^ „ 



"for 



shown in 



Fig. 4. 19, this relation becomes rather com- 
plex. However, consideration of this de- 
pendence radically changes the character of 
the conditions necessary for the occurrence 
of flutter. 

Figure 4-20 gives the values of critical 
flutter rpm at different magnitudes of the 
oscillatory blade motion in the feathering 
hinge cpfor> calculated with consideration of 
this nonlinear dependence as applied to flap- 
ping flutter. The calculation was made only 
for one value of blade balancing and different 
amplitudes of forced motion in the feathering 
hinge cpf o r • 

The curves plotted in Fig. 4. 20 permit a 
number of interesting conclusions. 



First of all, it follows from these curves that, in the presence of forced 
motion in the feathering hinge, flutter occurs at certain revolutions of the . 
rotor and its appearance is not due to the effect of any extraneous influence in 
the form of some initial impetus. In this case, the rpm of flutter onset is 
smaller, the greater the amplitude of forced motion in the feathering hinge cpfor » 
This fact is responsible for the dependence of the critical flutter rpn in /381 
flight on all parameters of the flight regime that determine the amplitude of 
cpfor* 3^<i primarily on the helicopter balancing and the flying speed. In ground 
tests, this leads to dependence of the critical rpm on the position of the con- 
trol stick. 



A second important characteristic of flutter, following from the curves 



411 



(see Fig. 4. 20), is the appearance of two different types of flutter which differ 
by the character of the increase in vibration amplitude upon any change in rotor 
rpn. 



Upon an increase in rpn to values 
flutter win set in with an . amplitude 




Fig. 4. 20 Variation in Critical 
Flutter Rpm with Vibration Ampli- 
tude in the Feathering Hinge cp, j^, 
at Different Magnitudes of Forced 
Motion. 



"hard flutter". 



corresponding to the points ai, ag, as, 
smoothly increasing with increasing rotor 
rpm. If, after the occurrence of 
such oscillations, which are usually 
called "soft flutter", the rotor 
rpm remains unchanged, then their 
amplitude will remain constant for 
as long as desired. 

Oscillations of this type have 
been repeatedly observed in ground 
and flight studies of flutter in 
helicopters. A decrease in rotor 
rpn after the occurrence of "soft 
flutter" leads to cessation of 
oscillations at the same rpm at 
which flutter began. 

Upon an increase in rotor rpm 
to values determined by the points 
bi and h^, oscillations are generated 
whose amplitude increases in time 
without an increase in rotor rpm. 
Oscillations of this type are called 



Probably, the limiting values of the vibration amplitudes obtainable in 
this case are determined by the nonlinear nature of the change in aerodynamic 
forces relative to the angle of attack. This branch of the curve in Fig. 4.20 is 
shown approximately by a dashed line. 

When "hard flutter" occurs during ground tests of a helicopter, the in- 
crease in blade vibrations can be stopped (to prevent an accident) only by a 
marked decrease in rotor rpm. The generation of such oscillations in flight may 
lead to serious consequences. 

A decrease in rotor rpn after the onset of "hard flutter" leads to cessa- 
tion of vibration at an rpm corresponding to the point k, which, as a rule, is 
smaller than the values corresponding to ax and ag . 

Thus, to stop "hard flutter" the rotor rpn should be decreased to values 
lower than those at which flutter began. 

At small amplitudes of forced motion in the axial hinge, the occurrence of 
"hard flutter" is possible only after some initial impetus, just as in the case 
when forced motion is absent. 



The rpm corresponding to the point nx should be considered the most 



7382 



412 



probable rpm for the start of "hard flutter" since, in this case, the magnitude 
of the necessary impetus is minimal. 

In calculating the critical rpm for the onset of flutter, corresponding to 
ai, as, as in Fig. 4. 20, additional simplifications can be made in eq.(3.12). 

As follows from Fig.4.19, when — ^ ^" <, 0.5, the value of the coefficient 
_ Vf or 

Afjy can be determined by the formula: 

If the frequency of forced motion is p, 5,, = mu) (m being the order of the 
harmonic of this motion with respect to rotor rpm), then we can write 

In this case, eq.(3.12) takes the following form: 

dfr--- ,'"''' . (3.14) 

The value of the equivalent moment of friction is here proportional to the 
rate of angular displacements and does not depend on the vibration amplitude of 
flutter cpf 1 u : 



^^i^-i^a;:^'^- (3.15) 



In other words, the moment of friction acting in the feathering hinge in 
the presence of forced motion in this hinge affects small oscillations of the 
blade in the same manner as a linear vibration damper, whose moment is propor- 
tional to the rate of relative displacement. This conclusion pertains not only 
to the feathering or axial hinge of the blade but is generally valid for all 
mechanisms with friction. 

It also follows from Fig. 4*20 that friction in the feathering hinge, even 
in the presence of forced motion, increases the critical flutter rpE in compari- 
son with the case where Mf, = and represents a useful factor from this point 
of view. Therefore, to improve the flutter characteristics of a rotor it is 
possible to use friction dampers in the feathering hinges. Of course, the use 
of such dampers is possible only when the helicopter has a sufficiently powerful 
and reliable booster control. 



413 



Section 4« Rotor . Flutter , yrith , Co nsideration of Coupling of 

Blade Vibrations through the Auto matic Pitch Control 

1. Forms of Rotor Flutter Observed in H elicop ter Experiments 



As mentioned above, the occurrence of flutter in a helicopter sets up vibra- 
tions of all rotor blades. These oscillations begin simultaneously despite the 
fact that the parameters of individual blades making up the rotor generally 
differ somewhat. Consequently, the simultaneous occurrence of flutter cannot be 

explained by the coincidence of the 
critical rpm of individual blades. 
Furthermore, it has been noted in almost 
all experiments on helicopters that the 
vibrations of all blades are strictly 
synchronized so that each advancing /383 
blade duplicates the motion of the re- 
treating blade with some lag in time. 
The vibration amplitudes of different 
blades increase simultaneously so that 
their magnitude on the different blades 
is approximately identical. Flutter of 
one individual blade of the rotor of a 
given helicopter is practically never 
observed. 



Plane of longitudinal 
control 




Swashplate of 

automatic 
pi tch control 



Fig. 4. 21 Diagram of Rotor Hub. 



This type of vibrations in flutter 
is ascribable primarily to the coupling 
of individual rotor blades through the 
automatic pitch control (Fig. 4. 21), 



The vibration mode of the rotor in 
which each advancing blade duplicates 
the motion of the retreating blade with some lag in time is usually called 
cyclic vibration mode. Such modes are very often encountered in studies of 
helicopter rotor vibrations. Therefore, they should be examined in greater de- 
tail. 



2. Analytical Expression for Cyclic Mode s__of Rotor^ Vibration 

For cyclic modes of flutter, distinguished by the fact that each advancing 
blade duplicates the motion of the retreating one, we can construct an analyti- 
cal expression determining the law of variation of the blade motion parameters 
in time. 



If we fix the point of reference in time such that for t = we have Pni=o = 
= 0, then this expression can be written in the following manner: 



pAr= Po«" sin (p^ - A^A'p J, 



(4.1) 



where 



Pig = flapping angle of the n-th blade; 



414 



go = angle determining the magnitude of blade deflection at the initial 
reference time, for t = 0; 
q = exponent determining the time rate of change of vibration amplitude; 
p == frequency of oscillations in flutter; 
Ai^B ~ phase shift of vibrations for two successive blades. 

Equation (4«l) is used for determining the motion of blades with ramibered 
N = 0, 1, 2, ..., Zt, - 1 (zb being the number of blades of the rotor). 

For a blade with N = z^, the law of change of variables should coincide 
with the law of motion of the blade having N = 0. Proceeding from this assump- 
tion, the phase shift Ai|rm should be a multiple of the azimuth angle between /384 
the blades, i.e., 

b 

At critical flutter rjm, for q = 0, the vibrations of all blades take place 
at constant and identical amplitude but with different vibration phases. The 
analytical expression for the law of change of variables at critical flutter 
rpm can be obtained by substituting eq.(4.2) into eq.(4«l) and setting q = 0: 



^N=%^m{^pt~Nm^y (4.3) 



It follows from eq.(4.3) that the vibration phase distribution for blades 
in cyclic modes may differ depending on the quantity m. The quantity m is 
called the order of the vibration mode and may vary from m = to m = z^, - 1. 
At m - z^, the vibration mode of the rotor, as follows from eq.(4.3), will coin- 
cide with the mode having the order m = 0. In like manner, for m > z^ all modes 
will be repeated.- Thus, for any rotor there can be Zj, different vibration modes 
corresponding to different orders m varying from m=Otom=Zt -1. 

Equations (4*1) and (4.3), derived above for determining the modes of rotor 
vibration, were constructed only for the variable Bn • However, all other para- 
meters characterizing blade motion vary in the same manner. Nevertheless, a 
certain vibration phase usually exists between them and the variable P^ • There- 
fore, in many cases it will be convenient to represent the law of change of 
variables in a complex form. With respect to the variable Pn > this can be 
written as 

p^=?„e"-'^^*'», (4.4) 

where 

It should be noted that, in forward flight of a helicopter, the blade exe- 
cutes also forced vibrations of cyclic modes since, in flight, each advancing 
blade duplicates the motion of the retreating blade. However, unlike vibrations 
in flutter, the forced blade vibrations in flight are strictly synchronized 

relative to the rotor rpm, so that each harmonic of vibrations of an order m 

415 



II 



will correspond to the vibration mode having the same order: 



?j,= ?^sinm(iot-Nf-y (4.5) 

Here, m corresponds to the order of the harmonic of forced vibrations. 



3. Cyclic Vibration Modes in Specific Gases a nd Con tr ol L oads 

The division of vibrations into cyclic modes is convenient in that only 
certain rotor control loops are loaded in the presence of each such mode. 
Therefore, the critical flutter rpm is determined by the rigidity of that con- 
trol loop which is loaded in the presence of the particular vibration mode under 
consideration. Of practical interest are only those rotor vibration modes /385 
that correspond to the smallest control rigidity and hence to the lowest critical 
flutter rpm. 

Let us study the manner of generation of cyclic vibration modes during 
flutter, in a specific case - for example - for a four-blade rotor. 



With a vibration mode of zero 
identical phases and load only the 



Position of coning 
axis daring 
flutter 



Position of coning 

axis before onset 

of flutter 

P 




Fig. 4. 22 Position of Coning 
Axis in Antiphase Flutter. 



usually called antiphase flutter. 



order (m = O), all four blades vibrate with 
collective pitch control. This form of 
flutter is called in-phase flutter. The con- 
trol rigidity referred to the axial hinge, 
and hence the critical in-phase flutter rpm, 
depend only on the rigidity of the collective 
pitch control loop. 

The vibration mode of the first order 
(m = 1), just as that of the third (m = 3), 
is of greatest interest since on helicopters 
it corresponds usually to the smallest con- 
trol rigidity and hence to the lowest values 
of critical flutter rpm. Vibrations of these 
modes are characterized by the fact that only 
the moment loading the lateral and longi- 
tudinal control loops is applied to the 
swashplate of the automatic pitch control. 

The opposite blades in modes of the 
first and third order oscillate in opposite 
phases. Therefore, this mode of flutter is 



The coning angle of the rotor in antiphase modes of flutter does not change. 
Therefore, the motion of the blades in these modes is conveniently character- 
ized by the motion of the coning axis (Fig. 4. 22). In vibrations of the first- 
order mode, the cone of the rotor is deflected relative to the original axis 
through an angle 3 and rotates about it with an angular velocity p^^ = Pfiu ~ '^ 
opposite to the rotor rotation. 

416 



Both the direction and magnitude of this angular velocity vary in the 
third-order mode: pa = Pfiu + ou. 

The vibration frequency of the variable forces in nonrotating parts of the 
control system, just as the vibration frequency of the fuselage during flutter, 
coincides in magnitude vri-th the angular velocity of rotation of the coning axis, 
which constitutes the basic difference between these modes. 

If the dynamic rigidity of the nonrotating parts of the control did not 
depend on the frequency of forces applied to it, then the values of the critical 
flutter rpm corresponding to modes of the first and third order would be identi- 
cal. However, in all experimental investigations of flutter, only vibrations of 
one of these modes, most often of the third-order (m = 3)> sre usually en- 
countered. In several cases, in particular when the control system includes 
inertia dampers, the first-order vibration (m = l) is observed in flutter. This 
is explained by the fact that the dynamic rigidity of the nonrotating part of 
the control, operated by the inertia inherent to its components, depends on the 
vibration frequency. Consequently, the hinge control rigidities corresponding 
to modes of the first and third order on a helicopter differ somewhat in magni- 
tude. Accordingly, the critical flutter rpm also differs. These considera- /386 
tions will be supplemented in Section 8.6. 



m=0 



m = / 



m=2 



N'o H\\2 



N-t 



N-2 



N-3 




n 




/ 


\. 






/ 


\ 





f 


\ 


/ 


\j 


1 


\ 


A 



^ 


n 

\ 2 


/I 


\ 




M 


L 


\ 





m= 


-3 




r 


V^ 


1 


\ 




\^ 


I 


\ 


M 


/ 


% 






/? 




^ 


^ 




^'J 


/ 


\ 




'\ 


J 


\ 


J 


f 



Fig. 4. 23 Vibration Phase Distribution in Different Modes 
of Flutter, for a Four-Blade Rotor at -^ = 1.75. 

Droring second-order vibration modes (m = 2), the opposite blades in each 
pair have an identical phase, and the phases of these pairs differ by half a 
period. The forces applied to the control during vibrations of this mode are 
locked on the swashplate of the automatic pitch control whose rigidity mainly 
determines the hinge control rigidity for this case. Since this rigidity is 
usually sufficiently high, the possibility of flutter with this mode, which is 
usually called the plate mode of flutter, is improbable within the operating rpm 
of the rotor. 



417 



The curves (Fig. 4.23) plotted on the basis of eq.(4.3) permit judging the 
character of the phase distribution by blades in all these modes for a four- 
blade rotor. 

4. Differential Equations of Rotor Flutter -with Consideration 
of Coupling of Blade Vibrations through the Automatic 
Pitch Control " 

Each rotor blade, during vibration, generates a moment acting on the blade 
pitch control system. The magnitude of this moment, taken relative to the 
feathering hinge axis of the hub, can be written in conformity with eqs.(2.19) as 

where N = 0, 1, 2, ..., z^ - 1 is the numeral of the blade. 

Here we have used the same notations as those given in Section 2 in deriv- 
ing the differential equations of flapping flutter of an isolated blade with 
axial flow past the rotor. Now, the number of equations has increased z,, times, 
i.e., as many times as there are blades in the rotor. 

If oscillations of individual rotor blades are in no way related and if /387 
each blade is attached to the hub as an isolated entity, then, after substitu- 
tion of 

^C = c«;,.('-Pyv+v-P;v). (^^y) 

into eq.(4.6), we obtain equations coinciding with eqs.(2.19). 

However, helicopters usually do not have such rotor designs. 

Generally, as a consequence of interference, the elastic angle of rotation 
of each blade in the feathering hinge 

yN=<fN-\-^^N (4.8) 

follows the deformations of individual rotor control loops, which in turn are 
determined by the totality of forces arising from all rotor blades. 

For the conventional rotor control system, this relation can be represented 
in the form 

Y/v=Y..^ +Y.sin^l^» + Y,cosi.L^>+Y:,^>, (4.9) 

where, as before, N = 0, 1, 2, . . ., Zi, - 1. 
Here, 



Yep - angle of rotation of the blade according to deformations of the 
collective pitch control; 



418 



Yx and Yz ~ amplitude values of the angles of blade twist as a consequence 
of defonnations of the lateral and longitudinal controls re- 
spectively; 
yI^p "= angle of rotation of the n-th rotor blade as a consequence of 
deformation of the swashplate of the automatic pitch control 
under the effect of forces completely balanced on the plate; 
it is assTjmed that, if all external forces are balanced on the 
swashplate, its deformation obeys the condition 

2:vr;=o; (4.10) 



N 



•a.p 



where t^^^ is the azimuth of the N-th blade reckoned from the plane of the 
longitudinal control with respect to the swashplate spider (see Fig. 4. 21); this 
azimuth is related with the blade azimuth by the expressioh 

where 

Qgt = angle of stagger of the rotor hub spider; 
Ate on ~ control angle of advance; 

I = blade angle of lag during rotation about the drag hinge; in 
Fig. 4- 21 the blades are shown in a position where 5=0; 

,(.V) ,(0) I »r2Jt 

If the rotor has three or less blades, then the quantity Ya!*? should be 
set equal to zero, since in this case there is no combination of forces which 
could be balanced completely on the swashplate. 

For a foixr-blade rotor, all values of yI^p are equal in modulus, i.e., when 
N = 0, 1, 2, 3 

lv^'l = --^- (4.12) 

This equality is not observed for a greater number of blades. /388 

If we introduce the concepts of rigidity of various control loops referred 
to the axial hinge of the blade, then the hinge moment acting on the blade due 
to the control can be expressed in terms of these rigidities and deformations 
of the corresponding control runs: 

^^^ol = <=.p yep + ^xY. sin ^%' + c,Y. cos i>l^' + c^^vl,^', ( 4. 13) 

where c^.p, c^, c^ , and Ca.p are the rigidities of the collective pitch control, 
lateral and longitudinal controls, and swashplate respectively, referred to the 
feathering hinge of the blade. 

a9 



The form of notation of eq.(4.13) assumes that the'rigidity c^.p remains 
constant regardless of the type of combinations of forces locked on the swash- 
plate. 

The values of the deformations of different control loops referred to the 
feathering hinge can be expressed in terms of the angles of rotation of indi- 
vidual blades Yn i^ "^^ represent eqs.(4»9) as a system of equations relative to 
the unknovms Yc,p» Yx > Yz > ^^^ Ya.p* The solution of the system (4«9) yields 
the following expressions for deformations of individual control loops: 



c.j,-—^yN, 



N 



When 



^ N 



N 



(4.14) 






,(N) 



For a four-blade rotor, Va.p can be detemiined by the formula 



'a.p 



■ COS nN 



2y. 



cos nN. 



(4.15) 



For a number of blades z^ > 4> "the quantity Ya'!p is determined by the 
expression 



^'"7==}- S cos^AT^YyvCos. 



2nm 



N. 



(4.16) 



m=2 



N 



Substituting eqs.(4.l4) and (4.16) into eq.(4.13), we obtain the expres- 
sions for the hinge moment from the control for a blade with the numeral N: 



fCA')— 1 






+ — c,cos^w2y^cos.)-^^) + 



N 



(4.17) 



+- 



N 



Substituting eq.(4.17) into eq.(4.6) and examining this equation to- /389 
gether with the first equation of the system (2.19), we obtain a system of dif- 
ferential equations of coupled blade vibrations at axial flow past the rotor: 



420 



where 






-i-^ Pi Sin ^(^n 2 (cpyv + -^n) sin ^.<^^) + 

* N 

+ 7" /'z cos 1-^^) 2 (^^ + ^^^^ ^os -I'L'^; + 

* . N 

+ 7-^^/ 2 cos-2^ArV(cp^+.p^)cos^^Ar=0. 



/n=2 



(4.18) 









(4.19) 



The system of equations (4.18) is a system of ordinaj?y differential equa- 
tions relative to the unknown functions P^ and cpfj , with periodic time-variant 
■coefficients together with the variable 



where 



^iN) = ,,t-Ni,%, 



(4.20) 



^%- 






5. Transformation, of Eqs.( 4.18) in Particular Cases where 
Cyclic Modes are the Solution of the Differential 
Equations of Rotor Flutter 

Let us check whether cyclic vibration modes are the solution to the dif- 
ferential equations (4.18) of rotor flutter written with consideration of 
coupling between blade vibrations through the swashplate. 

In the general case, the relation between variables in cyclic vibration 
modes of a rotor can be represented in the form 






(4.21) 
421 



where /390 

3o and cpo = angles of rotation of the blade with the numeral N = 

relative to the flapping and feathering hinges, which are 
P unknown functions of time; 
Mb ~ — 2 — ^ phase angle characterizing the vibration mode of the 
^ order m. 

Substituting eqs.(4«2l) into the differential equations (4. IS) and succes- 
sively varying the values of m from to Zi, - 1, we find that cyclic vibration 
modes are the solution to eqs.(4.18) only for values of m = (in-phase flutter) 
and z - 2 s m s 2 (plate mode of flutter). At these values, the differential 
equations (4. IS) are transformed into equations exactly coinciding with the 
equations of flutter of an isolated blade [eq.(2.19)]. Only the value of the 
frequency of natural vibrations of a blade in torsion entering the second equa- 
tion of system (2.19) becomes equal to 



during in-phase flutter (m = O) and 



P,„-Pc.p (4.22) 



Pf^ -Pa.p (4.23) 



during plate flutter (z^ - 2Sms 2), when all forces due to the blades close 
to the swashplate and the lateral and longitudinal controls and collective pitch 
control are not loaded. 

At the same time, cyclic vibration modes at m = 1 and m = z^ - 1 are the 
solution to the differential equations (4. IS) only in one particular case, when 
Cx = Cj . In this particular case, the differential equations (4. IS) are trans- 
formed into equations coinciding with eq.(2.19) for an isolated blade. Only the 
value of pt« in this case should be equal to 

P^^=Px=pz. (4.24) 

Thus, flutter of a rotor as a whole can be studied on the model of an iso- 
lated blade having a rigidity of attachment equal to the rigidity of the col- 
lective pitch control Co.p, with the cyclic pitch control c^ = c^ and swash- 
plate Ca. p taken separately. 

6. Rotor Flutter in the Pre^ence_ of_D ifferent Rigidity 
of Longitudinal and Lateral Cont rols 

To solve the differential equations (4. IS) in the case of c^ 7^ c^ , we can 
use the following method. Let us introduce the new variables: 



N 



N 



(4.25) 



422 



N 



Successively multiplying all terms of eqs.(4.18) by sin i|ri^p and by /391 

cos ij'i'llp and summing them with respect to N, we obtain a system of ordinary dif- 
ferential equations relative to the new variables of the following form: 

-f (02 (6,2 — r,2) ?y — 2 C,20)T1^ — rf,2U)'Tl,p = 0, 
Tip 4-flrnU)T]p + 2(Bep +rf,iU)2$p -|-Ci2Ti,.+ di2<or\f + 

-f 0)2 (6,2 — C,2) T]^ + 2C,2(0;*^ + C?,2«)% = 0, 
r„ + ar22<4+[(622-l)'»' + /'^]?f-2o/T)^-flr22«>^Tl,+ ^ (4*26) 

+ ^21^3 + ^2i">-.p + /'^'^'P — 2c2i(OTip — d2yy\? =0, 

% + rf22«% + [(^22 - 1 ) t"^ H- PI] nf + 2 0)if + d2^<0% + 
+ r2iT,p+Cl'2,(UTlp+jD2xTl3 -J- 2C2,U)';p^^C?2i'»^^?=0. 



The system of equations (4.26) can be solved by the conventional method 
for solving a system of differential equation's with constant coefficients. 

A similar method of reducing the problem to a system of equations with con- 
stant coefficients was used by Coleman and B.Ya.Zherebtsov in investigating the 
ground resonance of helicopters. 

We can show that the variables (4.25) can be expressed by the variables 
proposed by A.P.Rroskuryakov for investigating helicopter stability. 

In his works, A.P.Proskuryakov expressed the angle of rotation of the blade 
relative to the flapping hinge in the form 



P;v= Oo (0 + «i (0 cos .]j;v+ ^ (0 sin tl-w. 



(4.27) 



On alternately multiplying eq.(4.27) by cos ilr^ and sin 1];^ a^nd summing with 
respect to N, it will be found that 



a,(^)=— V,pArCOs<l<;v. 






N 



(4.28) 



i.e., the variables ai(t) and bi(t) virtually coincide with the variables Tig 
and I3 . 

423 



The use of the above method for solving equations of helicopter flutter at 
Cj f Cj and iJ. 7^ was proposed by L.N.Grodko. It was also used by V.D.Il'ichev 
for obtaining practical results. 

Section 5. Flapping Flutter of a Rotor in Forward Flight 

1. Preliminary Statements 

Experiments carried out on various helicopters showed that, in forward 
flight, flutter might set in earlier than under conditions of axial flow past the 
rotor, for example, in ground-testing. Therefore, a determination of the /392 
critical flutter rpm in flight is of appreciable practical interest. The basic 
problem requiring solution in this case is the degree to which the critical 
flutter rpm is lower in flight than on the ground. 

A variety of other important practical problems arises in this connection. 
For example, what parametric margin prior to flutter should be secured under 
ground-testing conditions so as to preclude the possibility of the occurrence of 
flutter in flight. 

All these problems can be solved if there is an opportunity to calculate 
flutter in forward flight, which permits determining, in particular, the de- 
pendence of critical rpm on the flying speed. 

Furthermore, difficulties arise in calculating the flutter in flight. 
These refer primarily to substantial complication of the differential equations 
describing blade Aribration. Therefore, in examining the problem, one should 
begin with these. 

2. Differential Equations of Blade Oscillations 
in Forward Flight 

The differential equations of torsional and flapping vibrations of a blade 
in forward flight are derived in the same manner as for the regime with axial 
flow past the rotor. Only the values of the relative velocities of the stream 
flowing past the profile should be calculated with consideration of the addition 
term due to forward velocity. These velocities can be written in the form 

i/^ = cD;?(7+(isin<)<), j 

, V cos 01 
^here V' = —^ 

Substituting eqs.(5.l) into eqs.(l.6) and then eq.(l.6) into eqs.(2.3) 
and (2.11), we obtain the differential equations of blade vi.br5.tion in forward 
flight: 



424 



P + Ki + l**!, sin<l.)u,p+^l+|x&|Jcos<ji4--L(J,2ft^si^2<l>^u)2p- 

1 1 ^ - _ 
^ P-^^ii (1 - cos 2^) (o2(p = o>2yo J ft?X (r+ [J, sin <|>) dr, 

-* 

- cos 24,)j a)2 + ;,^2 J <p + C2,^'+(rf2i -(ifti^ sin ^)(op + 

4- ([^2, - (iftlj cos <}- - -i- (1,2 6^1 sin 2i}.l U)2 + y.;7^2 j p ^ ^2 9 _^ 

+ 7^ J ^mA?^,„^ rfr - (o2 ^ J 6^ X (F4- [X Si n ^) cfr. 



Here, the coefficients Cis, cg^ , d^ , dgi , dgg , bgi , bgs are the same 
as in eqs.(2.14); furthermore, we introduce the following additional coeffi- 
cients: 





(5.2) 



/393 







"22 



T^^/irl*"^^'-- 



(5.3) 



Assuming that the particular solution B'^ and cp"*^ of eqs.(5.2) is found, we 
set, as before (see Sect. 2. 4) 

where B^ and cpa are the angles of deflection of the blade from a position 



425 



corresponding to its steady motion, determinable by the particular solution. 

Substituting these expressions into eqs.(5.2), we obtain the follovdng dif- 
ferential equations of disturbed motion of the blade in forward flight: 



? + ('^n + P'*!! sin <!)) (Op + ( 1 + |j.*}j cos ([) ^- — (i.26| I sin 2(1)) o)2p -J- 

+ ^12? + K2 + t^of !2 sin <1)) (o<p + 

+ [*i2 — i-[^6i,sinil>--l-p.26|{(l-cos2.1,)lu,2tp=0, 

9 + (0^22 + H'C'^z sin '^) 0)9 + jr^jj + -i- [xftij sin <}- + 

-r-f l^'ftlHl-cos2<}.)]<o2+;72 j9 + C2iP+(rf2i-pL6>,sin<i;)(Bp+ 



f { 62,-ix6i2COS(l>-^[*2*^>sin2.;-l«)2+x/7^2 h=0. 



(5.4) 



Here, the subscripts of the variables p^ and cp^ , which indicate that they 
refer to disturbed motion, are omitted for simplicity. 



3. Solution of Differential Equatio ns /394 

Equations (5*4) represent a system of differential equations with periodic 
coefficients. The solution of such a system can be -written in the form 






(5.5) 



where the functions Tg and T„ determine the content of the harmonic components 
of blade vibration in flutter. 

These functions can be written as 



^P = 2 (Pn cos rt^ + P„ sin «<)-); 

n 

^f = 2 (?« cos «<1< + 9„ sin raip) , 



(5.6) 



where n = 1, 2, 3, ... are constant coefficients determining the order of the 
corresponding harmonics. 

The critical flutter rpm in this case can be determined if eqs.(5.5), with 
consideration of eqs.(5.6), are substituted into the differential equations (5.4) 
and if the coefficients of like harmonic components are equated. This operation 
results in the formation of_a system of algebraic equations relative to the un- 
known coefficients Po , 9o > 3n » Pn » 9n » and cp„ . To solve this system, it is 
necessary to determine the roots of the characteristic equation whose order 

426 



depends on the number of harmonic components n retained in the solution. 

The solution of eqs.(5.4), with consideration of the harmonics, greatly 
complicates the calculation and at the same time - at the values of (j, < 0.4 
actually used - introduces no essential refinements into the calculation results. 
Therefore, in practical calculations we usually employ either the approximate 
method without consideration of the harmonic components or else the method of 
calculation with numerical integration of the equations of blade motion with 
respect to time. One of the versions of this method will be given in Section 7 
of this Chapter. 

4. Determination of Critical Flutter Rpm without Consideration 
of Harmonic Components of Blade Motion 

If the effect of harmonic components on the critical rpm is disregarded, 
the calculation of flutter in forward flight is no more complex than under con- 
ditions of axial flow past the rotor. An approximate solution, neglecting the 
effect of harmonic components can be obtained, if the periodic coefficients in 
the differential equations (5.4) are omitted"'. In this case, the forward flying 
speed is taken into account by introducing, into eqs.(l.6), the constant part of 
the functions depending on Ux . For this it suffices to set 

U2^^^2/^2fr2-^l-^-^y j (5.7) 

Then, the system of differential equations of disturbed motion can be /395 
written in the following manner: 



'f + '^22""f + [(jf +">') + *22 (l - Y t^'*2*2 ) »' 1 f + 



(5.8) 



where 



^^2 = 



^12 
II 



*22 



For a blade of rectangular planform, the coefficient bfg can be considered 
as approximately equal to -2. The coefficient bgg is small in magnitude and has 
no substantial effect on the results. 

Equations (5.8) differ from eqs.(2.19) for a regime with axial flow past 

the rotor only by terms of the type of f 1 — • \i^hfs)' This permits determining 

* This method was proposed by V.D.Il'ichev. 

427 






the critical flutter rpm in forward flight by eq.(2.27); -however, in the expres- 
sions of certain coefficients of eq.(2.24) entering this formula there appears 

an additional term of the type f 1 — [i^'b^s)' 

— *12C2l[l — — (^^^^2) — ^=^12 0^21— ^12^21], 



(5.9) 



Thus, disregarding all harmonic components of blade motion, the problem of 
determining the critical flutter rpm in' forward flight .can be reduced to solving 
the system of differential equations (5.8) with constant coefficients. 



5. Effect of Flying Speed on Critical Flutter Rpm 

/ 1 2 ->'- 
The effect of flying speed, definable by the term II — • p, bi- 



m 



eqs.(5.8) proves to be quite weak. Figure 4.24 shows the dependence of the 
critical rpm on the flying speed, determined by the value of p,, for three dif- 
ferent values of blade balancing. 

If follows from the graph (see Fig. 4. 24) that the critical 'flutter rpm /396 

drops by about 5 - 10^ with an increase in 
flying speed to values of |j, = 0,25 - 0.3. 

In experiments carried out on heli- 
copters, the effect of speed is somewhat 
stronger. This can be explained by the ef- 
fect of the following factors: 

It is shown above in Section 3 that, 
for small blade oscillations during flutter, 
the axial hinge with friction can be re- 
garded as a linear damper whose efficiency 
is smaller, the higher the angular velocity 
of relative displacements in this hinge 
during forced vibrations of the blade. 
Therefore, the critical flutter rpm in 
flight decreases with increasing relative 



'*/lu 














/ 




















1 


f 




















' 


^-F./r^*% 




nt 


■-^ 






x,„'25'/. 


^ 


















-^ 




- 


"" 






^m-^n 




— 




^ 


;;;^ 






































— 1 





0.Z 



«« 



Cf 



as 



Fig. 4. 24 Critical Flutter Rpm 
as a Function of Flying Speed. 



428 



displacements in the axial hinge and hence with flying speed, since relative 
displacements usually increase mth speed. Hence it follows that all factors on 
which the helicopter balancing depends may affect the flutter, since balancing 
determines the vibration amplitude in the axial hinge with respect to the first 
harmonic of rotor rpm. 

Displacements of the blade in the axial hinge, with harmonics higher than 
the first, may also have a strong effect. These harmonic components usually 
have smaller amplitudes of displacement but relatively high angular velocity, 
leading to an appreciable reduction of the effectiveness of the damping action 
of dry friction in the axial hinge of the blade. 

Thus, in many cases the severe drop in critical flutter rpm in forward 
flight is explained by a decrease in the damping action of friction in the axial 
hinge . 

A no less important factor capable of substantially influencing critical 
revolutions of flutter is the variation in the aerodynamic characteristics of 
the blade profile in connection with fluctuations of the value of the Mach number 
under forward flight conditions. As mentioned above, a change in M in the range 
from 0.5 to 0.9 causes a marked change in the aerodynamic characteristics and, 
what is especially important for flutter, a distinct shift in the position of 
the profile focus. 

Only the method employing numerical integration of the differential equa- 
tions of blade motion with respect to time (see Sect. 7) permits taking into 
account these factors with sufficient accuracy. 

Section 6. Calculation of Flutter with Consideration of 
Bending and Torsion of the Blade 

1. Bend ing and_Torsion of Blade during Flutter 

It was pointed out above that, in the overwhelming majority of cases, vi- 
brations of the blade as a solid body predominate in the mode of blade vibration 
in the flapping plane during flutter. The blade executes these oscillations, 
rotating about the flapping hinge. Torsional vibrations of the blade occur 
mainly as a consequence of its rotation about the feathering hinge. In this /397 
hinge, the blade rotates owing to the kinematic action of the swashplate of the 
automatic pitch control and flapping compensator as well as deformations of the 
control cables. Flexural and torsional deformations of the blade itself general- 
ly have no significant effect on the critical flutter rpm. Nevertheless, the 
flexural and torsional deformations of the blade during flutter of this type are 
usually quite pronounced. They lead to smaller displacements of the blade ele- 
ments in comparison with displacements during vibration of the blade as a solid 
body, but these displacements are of the same order. Therefore, it is impossible 
to neglect deformations of the blade itself or to show no interest in them. 

In individual cases, the flexural deformations of the blade increase and 
begin to have a noticeable effect on the critical flutter rpm. It is especially 
important to take into account blade bending in determining the effect of con- 

429 



centrated balancers installed on the blade to eliminate -flutter. 

Also known are individual cases where the blade during flutter executes 
flexural vibrations in which the share of the flapping mode is quite small. It 
should be emphasized that such cases are very rare. However, for jet helicopters 
with blade-tip engines, such flutter - usually called "bending flutter" - con- 
stitutes a serious danger. Subsection 8 of this Section will be devoted to an 
examination of this type of flutter. 

As stated above, the effect of torsional vibrations of the blade during 
flutter can be disregarded at a value of the coefficient a < 0.4 - 0.5 (see 
Sect. 1.4). In the remaining cases, in particular when the pitch control system 
has relatively great rigidity, blade torsion cannot be disregarded. This may 
result in a very large error.. 

However, most of the presently constructed helicopters have a coefficient 
a < 0.4* Therefore, in Subsection 6, we will specifically study flutter with 
consideration of bending but without consideration of torsional deformations of 
the blade. Such an approach leads to a considerable simplification of the dif- 
ferential equations. 

2. Determination of the Torque from^ ending Forc es on. the Blade_ 

In calculating torsional strains of a blade it is important what method is 
used for determining the torque due to bending forces on the blade. If the 
blade is bent in the flapping plane, then the force Q applied to the blade in 
the plane of rotation creates torque on the arm Ay relative to the section, at a 
radius r closer to its root (Fig. 4.25). Likewise, when the blade is bent in the 

plane of rotation a similar 
torque on the arm Ax is created 
by the force T acting in the 
flapping plane. 

In calculating the twisting 
moments due to bending forces 
on the blade, it is important to 
recall the fact that the com- 
ponents of the centrifugal 
forces relieving the blade in 
bending also participate in the 
generation of twisting moments. 
If we calculate only the torque 
due to external bending forces 
on the blade, the value will be 
much larger than the actual 
torque, just as the moment due 
only to the external forces bend- 
ing the blade will be many times greater than the bending moment in the blade 
section. 

Let us examine a blade element of length dr, bent in two mutually perpendic- 




Fig.4.25 Diagram of the Occurrence of 
Twisting Moments due to Bending Forces 
on the Blade. 



430 



ular planes (Fig. 4- 26), Equating to zero the sum of the moments of all external 
forces relative to the tangent to the blade axis in a section at the radius r /398 
and discarding all terms of higher orders of smallness relative to dr, we obtain 

-Mt„+Mf^^dM,^-M,^!'dr^M,y"dr=Q (6-1) 



or 



dr 



-.M^'-M^y". (6.2) 



If, for simplicity, we assume that the planes of maximum and minimum blade 
rigidity coincide with the planes of rotation and flapping, then, having set 



and 



we obtain 



V'=^ (6.3) 

My 

S'--^. (6.4) 



'-^-"Mir.-ii;)- f^-'' 



where I^ and ly are the elastic moments of inertia of the blade section during 
bending in the plane of rotation and flapping plane. 

Equation (6.5) was first proposed for calculations of a blade by V. N.Novak. 

It follows primarily from an examination of this formula that the torque 

BMtor 
— ^ per unit length due to the bending forces on the blade is always equal 



to zero if 



(6.6) 



i.e., if the rigidity of the blade in the plane of rotation and in the flapping 
plane is identical. 

Furthermore, by virtue of the smallness of the bending moments Mx and My 
(as a consequence of load relieving by centrifugal forces, these moments are by 
a factor of 8 - 12 less than the moments due to the external forces acting on 

the blade), the torque — ^°'' per unit length will be quite small in all cases 

even if I^ 7^ ly. This conclusion is highly important and results in a general 
approach to calculating torques and torsional deformations of a blade, as 
follows : 

In each section of the blade, we must determine the torque relative to /399 
the flexural axis of the blade in the examined section due to forces acting only 
in this section. Then, these local twisting moments should be summed with 

431 




Fig. 4. 26 Diagram of Loading the 
Blade Element with Stresses in Two 
Mutually Perpendicular Planes. 



respect to the blade length. Hence it 
follows in particular that the arms of 
the forces causing the twisting moments 
of the blade must remain constant re- 
gardless of whether or not the blade is 
bent. 

With regard to flutter calculations, 
it follows from this conclusion that the 
torque per unit length of the blade from 
centrifugal forces should be calculated 
by the formula 



dMt.. 



SL=m2 



dr 



= ia'^mray\ 



(6.7) 



rather than by the frequently used 
formulas of the type 



or 



-<sy^ y' ( mradr\ 



(6.8) 



(6.9) 



which holds true only for a blade with an infinitely great rigidity in the plane 
of rotation. 



3. Differential Equations of Binary Blade Vibration 

Binary blade vibrations in vacuum are examined in Section 5, Chaoter I of 
Vol.11. In studying binary vibrations in air, we must additionally take into 
account aerodynamic forces. Using the differential equations of flexural 
[eq.(l.9)] and torsional vibrations [eq.(5.6)] of a blade (see Chapt.I of /400 

Vol.Il) and supplementing these with inertia terms of the couple and with aero- 
dynamic forces expressed by eqs.(l.6) of this Chapter, we obtain a set of dif- 
ferential equations of torsional blade vibrations in air: 

my + [£//]" -lA^i/'l'-macp- 
- i- clQb \U\ + (^* - ^o) U!f - Uy]^ = 0; 



/mT-[Gr«']' + <oV„cp -l-^e6^f/? + 



+ \clQby [f^'?+(-f- *-Jro)f^?-^i]-<«''«»'-i/' -'«=i/=0. 



(6.10) 



432 



These equations are -written in a form pertaining only to disturbed motion 
of the blade. The particular solution describing undisturbed steady motion of 
a blade will not be discussed here. 

In eqs.(6.10), we use the following designations: 

y = displacement of the blade element in the flapping plane during 

disturbed motion of the blade; 
cp = angle of rotation of the blade element in the same motion; 
m = mass of the blade element per unit length; 
I„ = moment of inertia of the blade element per unit length relative 

to the feathering hinge axis; 
GT = torsional rigidity of the blade; 
N = centrifugal force in the blade section: 



A^ = (o2 f mrdr; 



a = distance from the center of gravity of the section to the feather- 
ing hinge axis, with the direction from this axis to the trailing 
edge of the blade considered positive; 
a^ = distance from the profile aerodynamic center to the feathering 
hinge axis. 

Differentiation with respect to the blade radius is denoted by a prime and 
with respect to time by a dot. 

To solve this set of equations it is convenient to change from the variable 
cp, which determines the total angle of rotation of the blade element in dis- 
turbed motion, to the variable ■& representing only the elastic angle of rotation 
of the blade and correlated with cp by the relation 

cp = & — xi/o. 

where 

Jq = angle of rotation of the blade in the flapping hinge; 
H = flapping compensator. 

Let us substitute the expression for the angle cp into the differential 
equations of binary blade vibration [eq.(6.10)]. This makes it possible to /401 
rewrite them in a form more convenient for further transformation: 



my^[EIy'T - [Ny-\ + mo^y, + 

+ -1 c^qb^ [fy^^o +(t '' " -^o) ^^"] " 



(6.11) 



433 



— may — ijy'o — w^I„xy'o ^ Qb^Uxy'o — 



In the presence of a horizontal flying speed of the helicopter, the rela- 
tive velocity of the flow past the profile will be a periodic ftmction of time 
and radius. This velocity can be set approximately equal to the velocity Uxt 



6^=u)r+Vsin<Bf. 



(6.12) 



Therefore, eqs.(6.1l) represent a system of partial differential equations with 
coefficients periodically varying in time. 

When the flying speed of the helicopter V equals zero, the periodic coeffi- 
cients of the system (6.11) become constant, independent of time. 

For the examined type of rotors system, eq.(6.1l) has the following 
boundary conditions: 



Afo = [£/t/"]o=x(A/, + A//r). 



(6.13) 



where 



M© = bending moment in the blade root; 
Mt = twisting moment in the blade root; 
Mfr = moment of friction in the axial hinge of the hub; 
Ccon ~ rigidity of the control system; 

«?o = angle of rotation of the blade root due to deformations of the 
control system. 



4. Solution of Differential Equations 

The solution of the system of differential equations (6.11) can be obtained 
by using B.G.Galer kin's method. We set 






(6.U) 



where 



y^^^ and ■d^^^ = modes of the natural flexural and torsional vibrations of 
the blade in vacuum; 
6j and Yk ~ coefficients of flexural and torsional deformations of the 



434 



blade with respect to the j-th flexural and k-th torsional harmonic 
of natural vibration. 

The coefficients 6j and Yw aj^e certain functions of time. Since eqs.(6.1l) 
are differential equations vrith periodic coefficients, the coefficients 6^ /402 
and Yk should be functions of time of the type 

S^=8^e"(l+7') (6.15) 

where the function T determines the content of harmonic oscillations during 
flutter. 

If, as before in Section 5*4, we seek the solution with an accuracy limited 
only by the fundamental frequency and disregard the effect of harmonic components, 
then we can omit the periodic coefficients in eq.(6.1l). 

Applying B.G.Galerkin's method to this simplified system of equations, we 
obtain a system of ordinary differential equations relative to the variables 5^ 
and Yk • In matrix form, this system can be written as before (Sect. 2. 5) as the 
equation 

Here the variable X is the vector function with projections 6j and Yk > i.e.. 




^- " (6.17) 



while A, B, C, and D are rectangular matrices of the order z, where z is the sum 
of the number of flexural and torsional harmonics accounted for in the calcula- 
tion. 

Setting X = Xoe^ * in eq.(6.l6), we obtain a system of algebraic equations 
of the form 

\CX^ + D<.\^A + o,^B\X,=0. (6.18) 

Let us then equate the determinant of this system to zero. The resultant 
algebraic equation relative to the unknown parameter X is the characteristic 
equation of the system (6.16). The roots of this equation completely charac- 
terize the blade motion described by the system (6.11). 

To determine the boundaries of flutter, we should set A. = ip in the charac- 
teristic equation and find the corresponding values of cu and p. These values 
will determine the parameters of the limits of the flutter zone. 

435 



w 



An analysis of the results obtained from calculations shows that, in the 
general case, each combination of torsional and flexural harmonics of blade vi- 
brations may correspond to a zone of instability vjith oscillations having a mode 
in which the content of the harmonics of this combination predominate. However, 
with actually used blade parameters, a given flutter zone by no means corre- 
sponds to each combination of harmonics. Thus, the nvmfcer of flutter zones is 
always smaller than the number of combinations of flexural and torsional har- 
monics and can never be greater than the number of these combinations. 

For practical purposes, an important point is the direct dependence of the 
critical flutter rpm on the frequency of the natural vibrations of the torsional 
harmonic of the blade entering into the combination in question. Therefore, 
combinations involving only the first harmonic of torsional blade vibration /403 
give the lowest values of critical flutter rpm. All other combinations based on 
higher torsional harmonics of the blade are of no practical interest since the 
critical flutter rpm corresponding to these zones is always higher than the 
operating range of interest here. 

All forms of flutter, corresponding to combinations of different flexural 
harmonics of the blade with the first harmonic of torsional vibrations of the 
blade, will be called the principal modes of flutter. Below, we will be in- 
terested only in the principal vibration modes since these modes of flutter have 
the lowest critical rpm and therefore are the only onfes encountered in practice. 

5. Calculation of Flutter with Consideration of 
Three Degrees of Freedom 

To illustrate the above method, let us examine in greater detail the compu- 
tational formulas for the case where the vibration mode during flutter is repre- 
sented as combinations of the zero r and the first y flexural and first tor- 
sional harmonics. 

The matrices entering eq.(6.l6) will be of the third order in this case, 
and the vector function X will have only three projections: 



Ar= 




(6.19) 



The coefficients of the matrices A, B, C, and D will be referred, as above, 
to the values of the coefficients Ih.h* Li, and I^ standing for the higher 
derivative of the variables: 

/:, = J /JMr; 


R 

/i = f my'^dr. 



' 1- 





436 



Let us write out the expressions for the coefficients of the matrices: 
a) Inertia matrix C: 



C = 



where 



^U ^12 <^13 \ 
Cji C22 C23 I ' 
^^31 ^32 ^^33 / 



mar dr. 



"-'+^J 









mar dr, 



c„ = 



r22 = l, 



R R -< 

J 



r23 — 



1 



« « -1 

J 

<^3i = y- I may dr, 





R 



b) Damping coefficient matrix D: 



/flf,l ^12 rf,3 

( "21 ^^22 ^^23 
W3I '='32 "^33 . 



where 



11 2 «' / 



A.A 



Lo p J 



""—■ J-'-tl'^^lT-f)'*- 



(6.20) 



(6.21) 



/404 



(6.22) 



(6.23) 



437 



rf.3 = 



R 



Cinf — 



*22 



LO J 





^1 



23 2 " Z, 






'A-^V^r 



* / 



--^Q-^PoJ*^^»rfr-x?„»„^,„ 



c/,,=-L^_^ 



2 '«' /, 



(■/ 





Lo J 



2a, 

if dj. = , ^'' [see eq.(3.14)]. 

TrLxntpp r 



c) Stiffness matrix A: 



/O 0' 
A=io a^^ I. 



^0 Oct. 



where 



^33/ 



-r,^ =— ^ 

,2 

V 



«22 = /J 
«33 = /'o 



(6.23) 



/405 



(6.24) 



(6.25) 



Here, po, is the frequency of natural flexural vibrations of the first harmonic 
of a nonrotating blade. 



d) Centrifugal and aerodynamic stiffness matrix B: 



438 



i 



where 



*U *12 *13 \ 



^ — I b^i ^22 *23 
\ *3l *32 *33 



;■ 



*n=l+4 



bn = Y cje^^ Po [j ft^' rfr + -I- H.2/?2 j 6r rfrl , 
*2i = - — f /raord rfr-1- X r fjb dr-\- 



'1-0 J 

''''^~^ [t '''^^''^° J *"' ^'^ '^'' + Y t^'^' f *°/ »'''■] 

lo - J 

k R 

u 

R K -I 

r ftr^i/rfr + -1- ii.2/?2 r & J/ rfr , 

J 

R -I 

l-|x2/?2 Uyddr , 

J 

R R -1 

J 



— C o 

2 " /, 



i!'32=-Y'^;77[|*^'^^'^"+ 



*33=*+^<^;e 7j-fo 



(6.26) 



(6.27) 



A06 



Here, k = ^= — rp^jmrdr where P = y'. 

The characteristic equation for this case will have the following form: 



XMo + X5a) A, + k< (B,(«2 + fij) + X«(« («)2C, + Cj) + X2 (a)«D, + oi^Dj + D3) + 



(6.28) 



439 



Hence, 



+ X(o(cO*£, + U)2£2+£3) + <»M'«*^l+«''^2+^3) = 0. 



^ 1 == flr„5o + c„5i + flr,3/?o + cig/^i + 0^127^0 + <^i27'i ; 

j9, = 6„5o + ^1 l-S"! -1- '^I A + *13^0 + ^13^1 + <^13'^2 + 

^2 = CjjOs -|- C13/C3 "T ^12' g! 

Ci = 6„5, + rf„52 + C„54 + 6,3/?! + rf,3/?2 + C,3/?4 + 

+ *I27') + ^12^2 + ^127^4; 
C2 = rfl,53 + CiiSg + rf,3/?3 + <r,3/?5 + rfizT'g + Cj27"s; 

£>, = 6„5o + rf„54 + C„5, + *,3/?2 + rf,3/?4 + '^13''?6 + 

+ *12^2 + ^127^4 + ^li^e'^ 
£>2 = *n53 + flfll'^s + ^^n^y + ^3^3 + ^13/?S + Cl3^7 + 

+ *i27'3 + dn^s + Cl2T^•, 

Ei = buS, + duS,-\-b,,/^, + di,R, + b,2r, + d,2Te-^ 

E2=^bnSs^dnS^ + b,^R,^d,^R^^b,2T,-{-d,2T^■, 

Cz'^dnOfj, 

F, = bnSe + b,sf^, + b,2Te; 

f2=bnSj+bi3R.j-\-bi2Tj; 

•'3 = ^11*^8' 



(6.29) 



where 



■ C32U23' 



Sq = C22C33 C23C32; 

•J 1 = f'22"33 r ^33"22 <^23^32 ' 

•^2 = ''22^33 ~l" ^33*^22 4" d2^33 — £'23'^32 ~ 

63 =^ ^22*^33 ~r ^33'-22> 

•J 4 ^^^ ^22"a3 ~r b 33d 22 — b^^d^i ''32"23i 

•^s ^^ o-iid^ "T ^^33^22! 



■ ^32'- 23 " 



-rfo.,^; 



23"32> 



"J 6 — b<^^- 



- b23"32-< 



'7 -^ 022^33 "T <^33^22' 



57 

"J8^^^^22'^33' 

^0 = ^21*^32 — '^22''3i; 

/?, = ^21^/32 + ^32^21 — <^22^31 — ^aidii, 



R2 — "2\^: 



21'- 32 



- ^32^21 "I" d2\dsi -^ ^22^^31 — ''3l'^22 " 



■ Ofoorf. 



22"31' 



/?3 — Cl22^3\'< 

/?4 = 601^32 + ^32'j'21 — 622^31 ~ *31^22'. 

/^5= 022" 31' 

^6 ^^ ^21*32 — ^22^31 ! 



/4O7 



(6.30) 



hhO 



^7=— «22*3i; 






^0= ^23*^31 ~ ^21^33; 






' 1 = '^23"31 ~t" ^31" 23 ^21*^33 — <^33''21' 






T2'=l^23C3l + *3lC23 + '^23^31 — ^21^33 " 


" ^33*^21 ~ 


-d^.dsi. 


'3= — ^33*^21; 






T^ = ftjgC^a, + 631^23 — ^21^33 — 633^21 ; 






T^= —a^d 2u 






7'6=*23*31— ^21^33; 






7^7=— «33*21 







The roots of the characteristic equation (6.28) can be determined by means 
of any standard program available for digital computers of any type. Such a 
program can include the operation of computing the coefficients of the charac- 
teristic equation directly from the coefficients of eq.(6.l6). In this case, 
eqs.(6.29) need not be used. 

The values of the angular velocity cu corresponding to the limits of flutter 
can be obtained also directly if, in the characteristic equation (6.28), we set 
X = ip and equate to zero the real and imaginary parts of the equation separately. 

The equations thus obtained vri.ll have the following form: 



^iC"', yt7) = ;'Mo-/j''(»25, + fl2)+/'M'«''£>i+"^'£>2 + -03)- 

io(u,, p)=/7M, - p2(to2C, +C2) + (B'«f , +«,2£2 + £3 = 0. 



(6.31) 
(6.32) 



If, from the equation L2(uu, p) = 0, we determine p = f(u}) and substitute 
into the equation Li(ci), p) = 0, then the points of intersection of the obtained 
curve Li(u)) = with the abscissa will correspond to the limits of flutter. 



Calculation of Fluttgr__with Three Degrees of 
Fr eedom Disregarding Blade Torsion 



/AQ8 



All the formulas presented above are appreciably simplified if we assume 
that the rotor blade is absolutely rigid in torsion. It was noted above that 
this assumption is valid for all rotors for which the torsional rigidity of the 
blade is appreciably higher than the rigidity of the blade pitch control system. 
In this case, during torsional vibrations the blade elements rotate mainly as a 
consequence of deformations of the control system and, to a lesser degree, owing 
to deformations of the blade itself. 

Consideration of a variation in the angle of rotation of the blade with 
respect to length leads to a minor change of certain coefficients of eq.(6.18) 
[see eqs.(6.21), (6.23), and (6.2?)]. This is explained by the fact that the 
magnitudes of the integrals entering the expressions of these coefficients are 
determined mainly by the blade tip which is subject to large aerodynamic forces. 



while the change in the angle of rotation s9 over the length of only the blade 
tip is insignificant. Therefore, the assumption of constancy of the angles of 
rotation of the blade cross sections over its length, in many cases, will not 
lead to substantial errors. At the same time, this assumption appreciably 
simplifies all computations, since i? = 1 and there is no need to decompose the 
angle of rotation of each blade section into ?? and kj4 • 

The differential equations of motion for this case can be written in the 
following manner: 





-\-{^b — x^U'<!}- Uy\ — moy\dr — 

R 

— u)2 J mray'dr=0. 



(6.33) 



The variable cp here represents the total angle of rotation of the blade 
relative to the feathering hinge as a consequence of deformations of the control 
and as a result of the kinematic action of the flapping compensator. 

The solution to this system of equations, just as for the system (6.10), 
can be obtained by means of B.G.Galerkin's method, if we put 

J 

9 = 90. 



where cpo is a function only of time and does not depend on the blade radius. 

Let us write out the computational formulas for the case where the vibra- 
tion mode in the flapping plane is represented by means of only the zero r and 
the first y harmonics of the natural blade vibrations. In this case, the coef- 
ficients of the matrices entering the equation of the forai of eq.(6.l6) can /409 
be determined by the following expressions: 



a) Inertia coefficient matrix C: 



C= 



'U 



■-21 



'-I2 



'-22 
C32 



'^33 



]■ 



(6.34) 



A42 



where 



R 

<^i2= — - — ( murdr; 

R 

^21= — - — \ma rdr; 
'ah J 

^22=1; 

R 

<^23= —. — {tnoydr; 
fahJ 

R 

^32= ——\maydr; 



^33=1; 



b) Damping coefficient matrix D: 



vfhere 



D- 



rf,l rfi2 ^13 

"21 "22 ^23 
"31 "32 "33 J 



d,z=\cl~^\br^.ydr- 
2 ^A.A J 



"22 — •■ 



'a.* 











+rf/r; 



(6.35) 



(6.36) 



(6.37) 



ZWD 



443 



d3i = YCl^^br^ydr; 

I 


R 



c) Stiffness matrijc A: 



A = 





(221 ^22 ^23 

0a 



where 



33 J 






^»f ' 



«22=/il = 






«23='''Po;>,^; 
a33=<; 



d) Centrifugal and aerodynamic stiffness matrix B: 



B = 



6,1 6,2 

"21 ''22 "23 
*32 *33 J 



where 



*u = i; 



6,2= ^c" -^ 

2 */,., 






69, = \ ma rdr; 



*22=1 + -^^!-,-^ 



2 «'/, 



*23 = 



» 

= — ■ I mar'idr; 

R R 

[br-^ydr^^-^W {bydr 



h — — JL/." P 

C/09 C- — 

^^ 2 " /, 



(6.38) 



(6.39) 



(6.40) 



(6.41) 



444 



II 



PM'"^"'- 



b^=k- 



The characteristic equations for this case will have the same form as in All 
the preceding case [see eq.(6.28)]: 



Here, 



where 






•^0 — ^ii'->o~r'-i2' o> 

^2=^U'->3T'^12' 3I 

D^ = b,,S^^dnS^ + CnST\-d,^R^-\-by^T^-^d,^T^-^c,J-^; 
^2 = bnS, + ^,,57 + ^^13^7 + bnT, + ^,2^; 

^3 = '^ll'^8 + '^12^8; 
/=', = 6„56 + &,27'6; 
•^2 = ^11'^7 4~^12' 7! 



•-*0 ^ ^22^33 '•23''32> 

O 1 = C22W33 "T ^33"22 ^23^32 '^32"23> 

Sj = &22<^33 4" ^33^22 "I" ^22^33 — <^23"32 "23*^32 "32C23> 

•J 3 = '^22'-33 "T '^33^22 ^23^32' 

•^4 = ^22*^33 "f" ^33^22 — ^23^32 ~" £'32"23l 

"J 5 = <l22'*33 "T ^^33" 22 023'^32> 

•Sg = b^ib^s ''23''32> 

S-j = ffl22''33 "F ^33^22 '^23''32> 

08 = fl22^33> 

"o^^^2l'^32> 

"l = ^21^*32 "T" ^32"21 — <-22"31' 

^2^''2l'^32 I ^32^21 "r"2l"32 ~~ "22"3i; 



(6.42) 



(6.43) 



(6.44) 



445 



f\3 = 0,21^32> 

/(4= O21W32 T''32'*21 ''22"31» 
/<5 = <Z2t^32 — ^^22" 31' 
°6^^^ ^21^32' 
/<7^ 021^32! 

'0^^ ^21^331 

/ 1 ^ CjaOai — C21U23 — C33«21 > 

/ 2 = U2^2l — ^21^33 ^33^21 — '*2l'*33> 

73= — 021^^33 — ^33*^21; 

'4 ^= ^23^31, — ^21^*33 — "3Sp2l ! 

/ g = ^23^31 ~~ '^2l'^33 — '^33"'2l ; 

^6^^ ^21^33! 

y 7= — '^21 ^33 — '^33^21; 

/8= 021^33* 



Zog 



The values of the critical angular 
mined by simultaneous solution of two e 




30 Xc^ % 



Fig. 4. 27 Critical Flutter Rpm as 
a Function of Blade Balancing, for 
Two Values of its Flexural Rigidity. 



velocities of a given case are deter- 
quations obtained from eq.(6.42) if we 
set X = ip, as in the case of the blade 
elastic in torsion. 



7. Calculation Results 

To illustrate the effect of 
flexural rigidity of a blade, Fig. 4. 27 
gives the critical flutter rpm as a 
function of 5Eo, ^ for a blade of mass 
constant over its length and with 
balancing. The curves are plotted for 
two values of flexural rigidity of the 
blade. The degree of rigidity is 
characterized by the values of the fre- 
quency of natural bending vibrations 
of the flxst harmonic of a nonrotating 
blade poj^ . The cases investigated are 
those of blades with the usual magni- 
tude of flexural rigidity, at 
POi/Ptw ~ 0.3 (solid curve) and of 
Poi/Ptw ~ 3«0 which corresponds to a 
very rigid blade (broken curve) . 



The share of bending in the mode of blade vibrations during flutter can /413 
be estimated from the ratios 6j/6g plotted for a mmber of points on the same 
graph. The quantity 61 /6p is equal to the ratio of the blade tip deflection in 
bending relative to the shape of the first harmonic to the displacement of the 

446 



tip during vibration of the blade as a solid body ( shape of the zero harmonic) . 

It follows from these data that for a blade with constant mass and balancing 
over its length, consideration of flexural deformations with respect to the 
first harmonic does not greatly refine the calculation results. 



8. Bending Flutter 



The results presented above cannot be extended to all designs of rotor 
blades. In individual cases, vibrations with primarily bending of the blade 
occur during flutter. This type of flutter is usually called "bending flutter". 

In bending flutter, the blade vibrates in the flapping plane with a mode 
close to some harmonic of the natural -vLbration of the blade in bending and is 
twisted with respect to a mode close to that of the first harmonic of the natural 
vibrations in torsion. As already noted, flutter with modes of subsequent har- 
monics of natural vibrations of the blade in torsion is theoretically also pos- 
sible. However, the critical rpm of such flutter is several times greater than 

the maximum rotor rpm. 

The previously examined flapping 
flutter can be regarded as a particular 
case of bending flutter in which the 
blade vibrates with a mode close to that 
of the zero harmonic of natural vibra- 
tions of the blade in the flapping 
plane . 



p 


y 


X 

y 




Ps 






^ 
y 


too 


/ 

/ 






■». 


— ■ 


— 






\ 





60 (O 




Fig. 4. 28 Variation of the Real and 
Imaginary Parts of the Roots of the 
Characteristic Equation as a Func- 
tion of Rotor Rpm. 



For footnote see next page. 



To each harmonic of bending vibra- 
tions of the blade there corresponds a 
separate flutter zone in which the vi- 
brations are characterized by specific 
parameters inherent only to this zone. 
Blade vibrations with different modes of 
flutter may occur quite independently. 
The mode of flutter having the lowest 
critical rpm is practically the first 
to be detected. Most often, this form 
is the flapping mode of flutter. How- 
ever, we can mention a number of par- 
ticular cases in which the critical rpm 
of some bending mode of flutter proved 
to be below the critical rpm of the 
flapping mode. 

As an example, let us discuss 
flutter of a blade with tip loading. 
This case is of practical interest for 
jet helicopters with an engine installed 
at the blade tip*. 



hhl 



Figure 4»28 shows the change of the real and imaginary part of the roots 
of the characteristic equation (6.42) with respect to rotor rpm. The roots of 
the characteristic equation (6.42) were calculated for a blade with a tip load- 
ing approximately equal to the weight of the blade itself. 

Figure 4.28 indicates that, in this case, there are two flutter zones; the 
flutter zone appearing first relative to rotor rpm is distinguished by a vibra- 
tion mode having a high content of blade bending. Therefore, this zone is usual- 
ly called the zone of bending flutter. 

It is possible to trace the manner in which the zone with the vibration /414 
mode having an increasing share of bending with increasing tip loading begins to 
separate from the zone of flapping flutter as the blade-tip loading gradually 
increases. At certain loading, these zones may separate into two different 
flutter zones. 

Figure 4.29 shows the flutter zone at a relatively small tip loading, equal 
approximately to 1/5 of the blade weight. In this case, the characteristic form 
of the zone of flapping flutter is distorted and the second zone begins to sepa- 
rate from it. 



equal to 42^ of the blade weight f = 0.42] and approximately equal to the 



Figures 4*30 and 4*31 show the flutter zones for a blade with a tip loading 

blade weight ( = l.l). In the latter case, the flutter zone separates into 

two different zones of flapping and bending flutter. 

6, 
Figures 4.29, 4*30, and 4«31 give the values of 6 = -^ — characterizing 

the vibration mode on flutter and the quantities p representing the ratios of 
flutter frequency to rotor rpm: 

tflu 

It is of interest that the share of the flapping mode of vibration in /4l6 
the bending flutter remains rather large in all cases, whereas the share of 
bending in the flapping flutter may be almost completely absent in certain cases. 

It should be emphasized that, for blades with tip loading, the critical 
rpm of bending flutter is appreciably below the critical rpm of flapping flutter, 
and that there is a weak dependence of critical rpm on the blade balancing. 
This fact greatly complicates the problem of developing blades for jet heli- 
copters. 



* The results of the calculations given here (in Subsects.7 and 8) were obtained 
by. V.M.Pchelkin. 

448 



/415 



W i/szc 




Fig. 4- 29 Flutter Zones with Blade Tip Loading Referred to 

n. 
Blade Weight ^° =0.20 



w t/sac 




Fig. 4*30 Flutter Zones vd.th Blade Tip Loading Referred to 

Gio _ 



Blade Weight 



= 0.42. 



449 



Vt/SiC 



eo 



w 



20 




F-r.i8 S=ij S=t.i9 s=v S=ws 6=-ijoi 

p^it.S p = 5 p=S.f5 p=SJ3 p=S.5l p=S.57 



20 



22 



2» 



2B 



28 



30 X,^ % 



Fig. 4.31 Flutter Zones vrith Blade Tip Loading Referred to 
Blade Weight -^^^^— = 1.1. 

9. Approximate Method of Determining the Mode of 
Bending Vibrations in Flutter 

If, in the first equation of the system (6.33), we discard terms of blade 
vibration damping as well as the small term mocp, then we can write this equation 
in the form 

my+\EIy"r-{Ny'\~clQbU\. ■ (6.45) 

Setting approximately U = our, we can represent the solution in the form 

, } (6.46) 



?=sin^^/, 



y=y^smp^^^ 

where pf^u is the vibration frequency of flutter. 

The calculations of bending flutter show that the frequency pf^^ can be 
approximately set equal to the^frequency of natural vibration of the blade in 
torsion and twist, i.e., Pfiu = Ptw* 



We assume that 



y^rT 



J 



ML 



450 



where 

6j = coefficients of deformations; 

y^^' = mode of .the j-th harmonic of the natiiral blade vibration. 

Substituting eq.(6.47) into eq.(6.45) and applying B .G .Galerkin' s method, 
we obtain expressions for determining the coefficients of deformation 6^ : 



»/ _ y;^/ 



(6.Z^) 



Here, V " / \ " / 

^, = JVVW; (6.49) 

^^= /n/? ' (6.50) 

where 

is called the equivalent mass of the blade during its Ad.bration relative to the 
shape of the j-th harmonic. 

Yj "^ mass characteristic of the blade during vibration relative to the 

same harmonic; 
Pj = frequency of the j-th harmonic of natural vibration of the blade in 

bending; for the zero harmonic y , we can set pj = ou. 

It follows from eq.(6.48) that the share of one or another harmonic of 
bending vibrations in the mode of flutter depends primarily on two parameters: 
relation of flutter frequency and frequency of natural vibration of the corre- 
sponding harmonic, and magnitude of the integral Aj . 

For example, the share of the flapping vibration mode ( j = O) in the mode 
of flutter is smaller, the higher the "harmonic" of flutter, i.e., the ratio of 
flutter frequency to rotor rpm. Here, we should note that since Aq is always 
greater than zero, i.e.. 



Ag^^b~r^dr>0, 



then the flapping mode will always be present in the vibration mode during 
flutter. This conclusion is rather important and indicates, in particular, that 
it would be incorrect to calculate flutter of some vibration mode without con- 
sideration of the flapping mode. 

The content of the first-harmonic natural vibration mode in bending in- 
creases as the frequencies of the natural vibrations in bending pi and the fre- 
quency of flutter pfiu close to the frequency of natural blade vibrations in 

451 



torsion are approached. 

However, in this case the magnitude of the integral A^ is very substantial: 



1 



The calculation of this integral shows that, for the vast majority of 
blades, this integral is close to zero so that the content of the first -harmonic 
natural vibration mode in the vibration mode with the frequency pn^ is quite /41B 
small. This explains the relatively rare appearance of bending flutter. 

A quite different picture arises when concentrated loads are mounted to the 
blade tip. The node of the shape of the first harmonic in this case shifts 
toward the blade tip and the absolute value of the integral A-^ begins to in- 
crease. Correspondingly, this causes an increase in the content of the first 
harmonic in the vibration mode with a frequency Pfm* 

Having assumed approximately that the vibration mode during flutter can be 
calculated in a form of eq.(6.47) where the coefficients are calculated by means 
of eq.(6.4B), we can develop a simplified calculation method for bending flutter. 

Section ?• General Method of Calculation of F lutter and 

Bending Stresses in the Rotor Blade during Flight 

1. Calculation Method and its Possibilities 

All methods presented above for the calculation of flutter were based on 
a number of assumptions which, in many cases, it would be desirable to discard. 
These assumptions include the following: 

1) In the calculation of aerodynamic forces, the nonlinear dependence of 
the aerodynamic coefficients on the profile angle of attack was disregarded. 
Consideration of this dependence may have a substantial effect on the critical 
rpm and especially on the character of amplitude build-up of oscillations in 
flutter. 

2) In calculating the aerodynamic forces under conditions of forward flight, 
the flow compressibility was accounted for by introducing only values of Cy and 
Xf averaged with respect to the rotor azimuth. Under conditions of forward 
flight these quantities periodically change with respect to rotor azimuth, which 
may have a noticeable effect on the critical flutter rpm. 

3) Consideration of the forces of friction in the feathering hinge, which 
- as is known - have a strong effect on the critical flutter rpm, was quite 
arbitrarily done, by linearization of these forces. 

In this Section, we will derive a method for calculating the bending and 
twisting (binary) blade vibrations of a helicopter in flight, which permits dis- 
carding these assumptions. This method makes it possible to determine the 

452 



bending stresses acting on the blade in the absence of rotor flutter and at 
stable blade vibrations. If flutter is possible in the operating regime of the 
rotor under consideration, then calculation by this method permits determining 
the process of divergent blade vibrations and thus investigating the phenomenon 
of flutter. 

The calculation method is based on the approximate solution of differential 
equations of blade vibration. In this case, B.G.Galerkin's method is used for 
determining the form of blade deformations at some instant of time, while the 
method of numerical integration of differential equations is applied for deter- 
mining the overall process of blade motion with respect to time. B.G.Galerkin^s 
method permits transforming the system of partial differential equations into a 
system of ordinary differential equations and to use numerical integration for 
solving this transformed system. 

As applied to stress analysis, the method permits accounting for torsional 
deformations of the blade in calculating the bending stresses in the flapping 
plane. Under the effect of constant and variable external forces in flight, the 
helicopter blade is twisted through some angle ?? which. is time-variant and /Z)l-9 
differs with respect to blade length. Torsional deformations of the blade 
change the angle of attack of its sections, which in turn leads to the genera- 
tion of additional constant and variable aerodynamic forces. These auxiliary 
forces must be taken into consideration when calculating the bending stresses of 
the blade. If this is not done, good agreement between calculation and experi- 
mental data is quite impossible. 

When applied to flutter calculations, the proposed method is not too con- 
venient in practical application, since it does not permit an exact numerical 
determination of the parameters characterizing the limit of flutter. The flutter 
limit can be established only in first approximation by visual inspection of 
ciorves describing the blade motion for parameters close to this limit; similar- 
ly, it is impossible to determine, with the required accuracy, the margins of 
flutter based on parameters used in practice for evaluating the rotor from the 
safety angle. The described method basically permits only a determination 
whether or not flutter occurs in the flight regime under consideration and a de- 
scription of its evolution. 

Nevertheless, the method has a number of important advantages in comparison 
with methods that use the roots of the characteristic equation and generally in- 
vestigate flutter only in a linear array. It is difficult to imagine any other 
method which would permit such a complete and accurate consideration of all non- 
linear dependences, both in the magnitudes of aerodynamic forces and in deter- 
mining friction forces, as is offered by this method in combination with numeri- 
cal integration of the equations with respect to time. Consideration of these 
dependences is highly important for flutter calculations. Therefore, it is 
preferably used in control tests and check calculations, after determining the 
flutter parameters by means of the roots of the characteristic equation. 

Of great importance lor practical use is the fact that this method, without 
excessive complication of the calculation, permits considering the elastic 
couple between blades through the automatic pitch control, even at different 
rigidity of the longitudinal and lateral controls. Without consideration of 

453 



this couple, a calculation of torsional deformations of the blade cannot lay 
claim to accuracy. 

2. Basic Assumpti ons and Suggestions 

To derive the differential equations of motion of the blade, let us examine 
the conventional type of rotor with individual hinge attachment of each blade 
to the hub and vrLth control through the swashplate. In determining the angles 
of twist of the blades as a consequence of deformation of the control system, 
we will consider that the rigidity of the longitudinal and lateral control loops 
differ. We will consider deformations of all control loops of both cyclic and 
collective pitch control including deformation of the swashplate, which is 
necessary when external forces generated by the rotor blades are locked on the 
plate. 

The motion of an individual rotor blade will be considered to consist of 
flapping and bending vibrations in the thrust plane and of torsional vibrations, 
both due to deformation of the blade and of the control system and to the kine- 
matic action of the swashplate and flapping compensator. As above, we will dis- 
regard blade vibrations in the plane of rotation. 

With respect to blade design, let us use the following stipulations: Let 
us consider that the flexural axis of the blade is rectilinear and coincides 
with the feathering hinge axis. The plane of least rigidity of the blade will 
be assumed to coincide with the flapping plane, i.e., with the plane going 
through the axis of rotation of the rotor and perpendicular to the axis of the 
flapping hinge. The flexural deformations of the blade will be determined in /420 
this plane. 

The rotor blade will be considered as a beam with the parameters continu- 
ously distributed over its length. 

3. Differential Equations 

With consideration of the above stipulations, the differential equations of 
blade vibration can be written in the following form: 

where 

y = displacement of points of the elastic axis of the blade relative 

to the plane of rotation of the rotors; 
9 = angle between the profile chord and plane of rotation of the rotor; 
m = mass of the blade per unit length; 
I„ = moment of inertia of the blade per unit length relative to its 

flexural axis; 
EI = flexural rigidity of the blade; 
GTtw = torsional rigidity or twist of the blade; 

454 



a - distaaice from the flexural axis of the blade to the centers of 

gravity of its elements, with the shift of the e.g. toward the trail- 
ing edge of the blade considered as positive; 

u) = angular velocity of rotation of the rotor; 

r = distance from the axis of rotation to the examined blade element; 

N = centrifugal force in the blade section: 



N=(iy^^ mrdr; 



T = aerodynamic load per unit length in the flapping plane; 
Sitae r ~ aerodynamic torque per unit length relative to the flexural axis. 

The method of determining the aerodynamic loads will be described in Sub- 
section 6. 

The dots in eqs.(7.l) denote differentiation with. respect to time and the 
primes, with respect to the blade radius. In differentiating the function 9 
with respect to the radius we should not introduce the geometric twist of the 
blade into the value of cd', assuming that cp' = ??' where ■& is the elastic angle 
of twist of the blade. 

4. Boundary Condit ions of the Problem 

For the type of rotors discussed here, the boundary conditions in the blade 
root can be written in the form 



(7.2) 



N 



where 



Mo = bending moment in the blade root; 

Mt = external torque relative to the feathering hinge axis due to 

forces acting on the blade, with the pitching moment considered 
as positive; 
K = flapping compensator; 
mi, = mass of the helicopter without blades; /421 

Z/[EIy" ]o = sum of forces striking the helicopter hub from all rotor blades 
^ (the index N denotes the blade numeral) ; 

Mfj = moment of friction acting on the blade in the feathering hinge 
from the side of the rotor hub, with the pitching moment con- 
sidered as positive; 
c^q = equivalent rigidity of the control system reduced to the axial 
hinge of the hub (the method of determining this rigidity will 
be given in Subsect.5); 
Y = angle of rotation of the blade root in the axial hinge of the 
hub, as a consequence of deformations of the control system. 

455 



In deriving these boundary conditions, friction was taken into account only 
in the axial hinge of the hub loaded by centrifugal forces. Usually, we can 
disregard friction in the other hinges of the hub and of the control system. 

With a sufficient degree of accuracy, the second boundary condition of 
eq.(7«2) can be replaced by the condition 

j/o = 0. (7.3) 

5. Determination of Equivalent Rigidity of the Cont rol System 

To use the third boundary condition, we must determine the magnitude of the 
equivalent rigidity of the control system c^g . This value can be determined if 
the angles of twist Yn °f ^•ll ^b blades of the rotor in the axial hinge of the 
hub are known. 

The angle of rotation of the N-th blade of the rotor Yn is related with the 
deformations of the individual control loops by formulas derived previously [see 
eq.(4.9)]: 



Y^ = Y..,+Y.sin^<,^' + Y,cos4,L^' + Yi:!', 



(7.4) 



where N = 0, 1, 2, 3, •••» '^■b ~ !• 

Solution of the system (7-4) yields the following expressions for its un- 
known Yep, Yx, Yz, andYi'^p: 



^'■P ' z, ^j'^'^' 



Y^ = 



-J-^^N^^^f^l 



* N 



)iN)., 



Yw - Y..n - Y^ sin <)-(,^) - Y^ cos<]>(,^) 



a.p 



ta.p ■ 



{1.5) 



The magnitude of the hinge moment acting on the blade from the control can 
be expressed in terms of rigidity and deformations of the corresponding control 
cables 



Moon =% Yep +^.Y.sini.i^> + c,Y.cos^l.^;+ c,.^y1C?. 



(7.6) 



where c<s.p, c^, Cj, , and c^^p are the rigidities of the collective pitch control, 
lateral and longitudinal controls, and swashplate, respectively, reduced to the 
axial hinge of the blade. 

If we represent the magnitude of the hinge moment due to the control in /422 



456 



the form 



JVicon —^tq 'A^> f 7 7^ 



then the equivalent control rigidity can be deterniined by the formiila 

<'=<^..y, f.;,+c,f,sintL^'+r,YzC0s4.i^> + c^^Y'j;>, (7-8) 

where the vinculimi denotes that the given magnitude of twist pertains to the 
value Yn • 

6. Determination of Aerodynamic Forces 

To solve a system of differential equations (y.l), it is necessary to de- 
termine the aerodynamic forces and torque entering the equation. 

It is known that, during flow past the blade profile in flight, the angles 
of attack of its sections may vary within wide limits, even to the extent that - 
on the retreating blade - the flow passes over its root parts from the side of 
the trailing edge. Flow-separation conditions occur at the blade tip in certain 
regimes. At high flying speeds and at appreciable peripheral rate of rotation 
of the rotor, the effect of flow compressibility has a considerable influence on 
the magnitude of the aerodynamic forces. Therefore, a determination of aerody- 
namic forces acting on the helicopter blade should take into account the non- 
linear dependence of the aerodynamic coefficients on the angle of attack a and 
the Mach number. Correspondingly, the expressions for determining the aerody- 
namic forces should be written with consideration of the possibility of a wide 
change in the angles of attack. At the same time, we can make use of the gener- 
ally employed assumption of smallness of the displacements y and angles of rota- 
tion of the blade sections cp. Therefore, to determine the aerodynamic forces 
the following expressions can be used: 

(7.9) 

where 

Cy and Cx = aerodynamic lift and drag coefficients; 

m^ = torque coefficient of the profile, with Cy, c^ , and m^ deter- 
mined from the results of downwash exposure as a function of 
the section angle of attack a and M; 
p = air density; 

b = blade chord in the examined section; 
Xo = distance from the leading edge to the flexural axis of the 
blade; 
Ux and Uy = mutually perpendicular relative velocity components of the 
flow in a plane normal to the elastic axis of the blades, 
with Ux being parallel to the plane of rotation of the rotor 
and Uy perpendicular to Ujt ; 

457 



I ■ III I 11 II II 



U = total magnitude of the relative velocity of flow past a profile in 
a plane normal to the elastic ax±s of the blade. 

The magnitude of the relative velocity U can be determined in terms of /423 
its components 



where 



Here, 



6^^=o)r+Vcosa^sirnj)^; (7.II) 

Uy=wRl-y~V cosa^cos%?. (7.12) 

cbR = tip speed of the blade; 

V = flying speed of the helicopter; 
or^ = angle of attack of the helicopter rotor in the shaft axes, i.e., 

angle between direction of flight and plane of rotation of the 

rotor; 
lift = azimuth angle of the blade; 

3 = y' = angle of inclination of the elastic axis of the blade in the flap- 
ping plane; 
X = velocity of flow through the rotor referred to the peripheral 
blade tip speed ouR, with the direction of X coinciding with the 
axis of the rotor shaft; when the flow passes through the rotor 
from the bottom up, X is considered positive. 

The relative velocity of flow is determined by the formula 

X=(x/^a^ + ^.^^, (7.13) 

where 

V cos Ol 

Here, Vj„a is the induced part of the velocity of flow, also referred to cuR. 

The induced velocity Vmd is a variable with respect to the rotor disk area 
and to time. 

In a number of flight regimes, the variable part of the induced velocity 
increases so much as to lead to the occurrence of appreciable variable stresses 
in the blade (see Sect .8, Chapt.I of Vol.Il). To determine stresses in the 
blade with consideration of the variable field of induced velocities, it is sug- 
gested to use the calculation method which involves calculation of the induced 
velocities. If we limit ourselves to a consideration of only the constant com- 
ponent of the induced velocity, then its value can be determined from the formula 



Ct 



4 y^^i^+OcoJ^' 
458 



(7.14) 



where 



Here, 



voay and Xp^y = components of the induced velocity and flow-through ve- 
locity, constant with respect to the azimuth and average 
with respect to the radius of the blade; 
Cj = thrust coefficient of the rotor: 

C = ^/v»^ 

Trot ^ rotor thrust; 
F = rotor area. 



The angle of attack of the blade sections, needed to determine the aerody- 
namic coefficients, can be calculated as 

where 

cp = angle of setting of the blade profile; 

f = angle of inflow: /424 

<^^fan-'~. (7.17) 

'-'X 

The angle of setting cp is a variable with respect to both blade radius and 
time. It consists of two parts: 



where 



9=T,+», (7.18) 

1] = angle of rotation of the blade in the feathering hinge as a conse- 
quence of the kinematic action of the swashplate and the flapping 
compensator, including also the geometric twist of the blades; 

?? = angle of elastic twist of the blade, with the angle j? determined by 
solving the system of differential equations (7.I). 

The angle 11 is determined by the expression: 

Tl=9o + Av<,„-e,sin^^-92COS<l>^-rjJo. (7.19) 



Here, 



00 = angle of setting of some blade section taken as the point of 
reference at 3o = 0; this angle is usually called the "indi- 
cator" angle of setting since its value is often given on the 
instrument panel of the pilot; 
^gooB ~ geometric twist of the blade; 
9i and 02 = angles of cyclic pitch control caused by tilting of the swash- 
plate; 
Po ~ angle of rotation of the blade in the flapping hinge. 

The Mach niimber, also needed for determining the aerodynamic coefficients, 

459 



I. 
1 



is calculated by the formula 

M = 

a* 

where a^^ is "the speed of sound. 



^=— ■ (7.20) 



Thus, eqs.(7«l) together with eq.(7«9) make up a set of partial differential 
equations with coefficients representing complex nonlinear functions of vari- 
ables . 

7. Method of Solving the Differential Equations 

The method of solving eqs.(7«l) most convenient for practical use at the 
present state of the art of computer technology is the method of numerical inte- 
gration of the equations of blade motion with respect to time, in which the 
blade deformations are determined by B. G . Gal er kin's method. In the formulation 
of the problem adopted here, this method permits obtaining the most accvirate 
results. 

In determining the bending strain of a blade, it is natural to represent 
the solution by means of functions which are natural vibration modes of the 
hinged blade in vacuum. The peculiarities in the distribution of rigidity and 
mass characteristics over the blade length and the boundary conditions of the 
problem have already been covered by such functions. We set 

^=2)w^ (7.21) 

where 

y •'' = mode of the j-th harmonic of natural blade bending vibrations; 
6j = coefficients of blade deformation with respect to the j-th 
harmonic. 

In determining the torsional strain, certain difficulties are produced /42$ 
by the fact that the deformations of the' controls vary substantially, depending 
on the direction of the moment of friction in the axial feathering hinge and on 
forces generated at the swashplate by the totality of rotor blades. The relations 
between the twist of the blade root and of all its longitudinal sections also 
vary, depending on the conditions of the effect of these factors. To take this 
into account, we must introduce some additional variable into the calculation. 

Let us study this problem in greater detail. To determine torsional de- 
formations by the Galerkin method, just as in determining bending deformations, 
it is logical to use the modes of natural torsional vibrations of the blade in 
vacuum. Here we can use various systems of eigenf unctions, differing by the 
boundary conditions in the attachment of the blade at the root. 

The solution to eqs.(7.l) is simplest if we assign the blade twist by means 
of natural torsional vibration modes, determined for a blade represented as a 
beam with a fixed value of torsional stiffness at the point of attachment 
(Fig, 4. 32a). This method of solution is quite common in practice. However, 

460 



here the problem basically reduces to a calculation of the vibrations of an iso- 
lated blade, since the use of the indicated modes precludes the possibility of 
accounting for the elastic couple between the blades through the swashplate. 
The effect of the moment of friction in the axial hinge of the hub cannot be 
fully covered. Actually, the elastic tvd.st of the blade root is determined by 
the magnitudes of the moments M^oon acting on the control system; furthermore, 
the magnitude of these moments at known moments due to the blade M^, depends on 
the direction and magnitude of the moment of friction: 

AI<^j=A^(^)+yw<^). (7.22) 

Therefore, blade twist at the root, and consequently the connection between the 
twist of all sections of the blade length, are related with the magnitude of the 
moment of friction. This effect cannot be accounted for if the indicated con- 
nection between the twists is fixed by vibration modes used in the calculation. 

It follows from the foregoing that this calculation method should be con- 
sidered invalid as applied to real helicopters. It can be used only in indi- 
vidual - rarely encountered - particular cases. 

To take into account the couple between blades through the swashplate and 
the effect of the moment of friction in the feathering hinge, we could use a 
system of functions representing the modes of natural torsional vibrations of 
the blade in the form of a free beam unattached at the root (see Fig.4-33D). 
However, owing to the discrepancy of boundary conditions, the use of such func- 
tions might lead to a solution of only an approximate type. Actually, the modes 
of torsional deformations thus obtained will substantially differ from the real 
modes. This difference will be especially pronounced in twist of the root por- 
tions of the blade where, for a free beam, the torque diagram drops to zero. 

All these considerations necessitate applying a nonorthogonal system of 
functions to this problem, as shown in Fig. 4.32c. In this case, the twist of 
the blade can be represented in the form 

?=t1 + Yo+2y,»'*'. (7.23) 

where k = 1, 2, ... . 

Here, 

Yo = angle of twist of the blade as a consequence of deformation of /426 
the control system; 
^(k) = mutually orthogonal modes of natural torsional vibrations of a 
blade rigidly fixed at the root; 
Yt„= unknown coefficients of the torsional deformations of the blade. 

Thus, the blade twist is represented by a system of orthogonal functions 
3?^"^ , supplemented by a function #°^ =1 nonorthogonal to this system. 

Equation (7.23) can be written in the form 

461 



f = n + ^y^bW, 



(7.24) 



where k = 0, 1, 2. 

This form of representing the blade twist creates certain complications in 
the calculation, produced by the nonorthogonality of the functions d^ . Never- 
theless, we must put up with these complications in order to accoiint for all of 
the above highly important factors. 



Beam vith elastic 
attachment 



Beam with rigid 
attachment 




Fig. 4.32 Modes of Natural Torsional Vibrations of a 
Beam with Various Attachments. 



8. Transformation of Partial Differential Equations 
into Ordinary Differential Equations 

Having represented the solution of system of differential equations (7«l) 
in the form of eqs.(7.2l) and (7.24), let us apply the Galerkin method. For 
this, let us twice differentiate eqs.(7.2l) and (7.24) and substitute them, to- 
gether with their second derivatives, into eqs.(7.l). 



form: 



The second derivatives from eqs.(7.2l) and (7.24) will have the following 






(7.25) 



462 



TABLE 4.1 





j = 


7 = 1 


;--2 


y-3 


* = 


k^\ 


*^2 


y = 


y-i 


;-2 


y-=3 


*-0 


k=l 


k=^2 






h 


81 


Bj 


53 


Yo ■ 


Yi 


Y2 


»0 


»i 


5j 


83 


Yo 


Yi 


Y2 




/ = 


*oo 


*0I 


*02 


*03 


.Coo 


coi 


f02 


«00 


«Q1 


''02 


"03 


*00 


*01 


*01 


-^0 


/ = 1 


*I0 


ftn 


*I2 


*I3 


ClO 


'u 


C12 


"10 


"11 


"12 


"13 


*10 


*11 


*12 


=-4. 


1 = 2 


*20 


*21 


*22 


*23 


^20 


C21 


<^22 


"20 


"21 


"^22 


"23 


*20 


*21 


*22 


=^2 


/ = 3 


*30 


*31 


*32 


*33 


C30 


C31 


er,2 


«30 


"31 


"32 


"33 


*30 


*3. 


*32 


=^3 


* = 


foo 


«01 


C02 


«03 


/o 


/i 


/2 


*00 


*01 


*02 


*03 


'■>%+Clf 


<o2/, 


a.2/2 


= So + Afjr 


* = 1 


"lO 


cu 


«I2 


fl3 


/l 


ii 




*10 


*n 


*12 


*13 


c«2/, 


v?^. 




=Bi 


*=2 


C20 


Cj, 


''22 


^23 


^2 




L, 


*2I) 


*21 


*22 


*23 


...2/2 




v^i2 


=B, 



^ 

UJ 



4> 



We then multiply the first equation of the system (V.l) by y^"*^ and the /428 
second by d^^^ and integrate all terms with respect to the blade radius. The 
boundary conditions (7«2) should be accounted for in the integration. This op- 
eration transforms the system of partial differential equations into a system of 
ordinary differential equations relative to the new variables 6j and Ytw 

For practical purposes, it is highly important what number of variables 6j 
and Ytw^^ used in the calculation. Experience has shown that a sirfficiently 
complete answer can be obtained if the bending strains are represented by means 
of the first four harmonics of the natxoral blade vibration and the torsional 
strains by two or - in the extreme case - by three harmonics. Thus, the problem 
of bending and twisting vibrations of a helicopter blade can be solved with the 
use, in any case, of seven independent variables. We will restrict the further 
calculation to this number of variables. 

The system of ordinary differential equations obtained from application of 
the Galerkin method is written out in the form of a table (see Table. 4.1) • 

All equations of this system represent the sum of the products of certain 
constant coefficients and the unknown functions 6j and Ytw ^■'^d their second 
derivatives. In Table 4.1, the coefficients pertaining to one equation occupy 
one row. The known constants that do not change during the calculation are 
written out in the squares of the table. 

The independent variables 6j and Ytw ^^nd their second derivatives, entering 
simultaneously all equations of the system, are extended with respect to the 
vertical in a special row in the upper part of Table 4»1. The right-hand sides 
of the equations are extended in a special column next to the table of constants. 

The coefficients of the left-hand side of the equations of the system (see 
Table 4.1) are determined after calculating the modes and frequencies of the 
natural blade vibrations in bending and torsion. As stated above, in calculating 
the torsional frequencies a blade rigidly fixed at the root is used. 

A number of coefficients are determined directly during this calculation. 
This concerns primarily the frequencies of the natiu^al vibration of a rotating 
blade in bending pj and in torsion \^ , and also the coefficients into which the 
mass characteristics of the blade enter: 





ft 



(7.26) 



After calculating the modes and frequencies, we determine the coefficients 
into which simultaneously enter the data obtained from calculating the blade 
in bending and in torsion. These are the following coefficients: 



464 



A,^ =f my^'^y^ndr +x2pu)pa)/„_i_ ^pu) J moyU)clr + 



+ ;cp(/)jmai,(Orfr. 



a„^p) J mi/<'Vy)rfr+x2«.2%op^(/)/^+ 



_|_i 



L 



(7.27) 

7429 



The second, third, and fourth terms in these expressions are small in com- 
parison with the first and can be neglected. At i 71^ j, the first terms of 
eqs.(7.27) vanish by virtue of the orthogonality of the functions y^ ' ^ and j t 
and the coefficients kij and a^j can be assumed as approximately equal to zero. 



x.e.. 

Next, the coupling coefficients are determined whose value depends mainly 
on the blade balancing. 



At j = 1, we have 



R I 



(7.28) 



The terms on the right-hand side of the system of equations (see Table 4.1) 
are determined by means of the following expressions: 






4-(B2(6jSlntl.-|-G2COS<;.) fmoi/Orfr— ? morpo dr , 



■J 



5.= aJ?^r»<*'rfr-«)2/L 



J/, — 1 JJiatr 




(7.29) 



465 



p 

Here, I^ = Golk + J I.Acpg,,,.^^ ''^ dr. 





The first terms of eqs.(7.29) are those determining the value of the coeffi- 
cients Aj and B^ . The following terms are small and can be neglected. 

9. Determination of the Magnitude of the Moment of Fri ctlonin the 
Feathering Hinge of the Hub 

During the numerical integration of the equations (Table 4«l)» the magnitude 
of the moment of friction can be obtained from the values of the torsional de- 
formations of the blade determining the external torque in the feathering hinge 
and from the direction of blade rotation in this hinge. In so doing, the magni- 
tude of the moment of friction should be determined by a different method, de- 
pending on which is greater in absolute value: the external torque in the 
feathering hinge M^i or the maximum possible moment of friction Mf^ • 

The external torque in the feathering hinge is determined by the formula /430 

M,.= M^-M,„„=^ V,„vM<*>, (7.30) 

where k = 0, 1, 2, 3, • • • • 

Here, 

Ml, = hinge moment due to forces acting on the blade; 
Moon ^ ^eqYe "" moment relative to the feathering hinge due to the control 

system; in conformity with this notation the pitching moment 
due to the control is considered as positive just as in 
eqs.(7.6), (7.7), and (7-22); 
Vi^^'' = magnitude of the hinge moment in blade deformations with 

respect to the mode of the k-th harmonic of natural vibrations 
of the blade in torsion. 

Modes of natural vibration normalized in some manner, for example, by the 
quantity j?r ' = 1, will now be discussed.' Here, we asstmie that 

The magnitude of the maximum possible moment of friction K't^^ is usually 
determined experimentally in the laboratory. If the coefficient of friction in 
the bearing f is known, then this magnitude can be determined by the formula 

where 

Np = centrifugal force acting on the bearing of the axial hinge; 
r^e = radius of this bearing. 

If JMhi I < JMfr^l, then Mf^ = -Mhi . In this case, the blade in the feather- 
ing hinge does not turn, and cf^ = c^ =0. This condition permits determining 
immediately '^o and Yo • 

Zp66 



If iMhi 

} (7.31) 






10. Sequence of Performing the Calculation 

The system of differential equations (see Table 4.1) is written here in a 
form such that its solution is conveniently found by numerical integration with 
respect to time. During this integration, mainly the right-hand sides of the 
equations will change. All coefficients on the left-hand side of the equations 
remain unchanged during the calculation, with the exception of the coefficient 
Cgq whose magnitude is recalculated at each integration step. 

The numerical solution of the system (see Table 4«l) also represents the 
basic part of the method of calculating binary blade vibrations presented here. 

The calculation of blade vibration by this method is carried out in the 
following sequence: 

1) Calculate the modes and frequencies of natural blade vibrations in 
vacuum. For calculation by this method, it is necessary to determine the first 
four harmonics of flexural vibrations of the blade, including the so-called zero 
harmonic of vibration of the blade as a solid body, and the first two harmonics 
of the torsional vibrations of a blade rigidly fixed at the root. From re- 
sultant vibration modes, determine the constant coefficients of the system of /431 
differential equations (see Table 4«l)« In" the numerical integration of the 
equations, all these coefficients remain unchanged with the exception of the co- 
efficient Cgq whose determination is described in Subsection 5« 

2) Select the parameters of the flight regime p , u), M-, ffh , 9© > 9i » ^s i^^ 
which the bending and twisting vibrations must be calculated. 

Usually, these parameters are taken from an aerodynamic calculation of the 
rotor and from calculation of the balancing characteristics of the helicopter. 
However, another more natural method can be used. The calculation method pre- 
sented here can be used as a method of aerodynamic calculation and calculation 
of balancing, by adding a number of simple operations. The values of o-h and %q 
can be obtained from the calculation if the values of thrust and propulsive 
force of the rotor and the angles 0i and Gg necessary for fulfilling the flight 
regime are prescribed and if the moments on the hub necessary for balancing of 
the helicopter are determined. 

3) At the initial instant of time, which is usually related with the azi- 
muth angle tb ~ 0> assign arbitrary values of the variables and their first 
derivatives 5j , YtH» ^j ^^^ YtM* To account for the coupling between the blades 
through the swashplate, these values are assigned for all z,, blades of the rotor. 

4) Determine the magnitudes of the aerodynamic forces necessary for calcula- 
tion: 

467 






(7.32) 

J 

■where the value of T] is determined by differentiation of eq.(7.19)t 

7]= — (1)9, cos ijj^ + 0)62 sin ij)^— xPp. (7*33) 

Here, Po = ^ 6jp^^^ . 

5) From eqs.(7.10), (7.11), and (7. 12), determine the velocity of flow past 
the profile and its components, and derive the angles of attack of the sections 
from eq.(7.l6). Use eq.(7«20) for determining the Mach number. 

6) From the polars of the profile fed into the computer together with the 
initial data, determine the values of c^, Cy, and m^. After this, making use of 
eqs.(7«9), calculate the aerodynamic forces per unit length T and the torsional 
moments 311^,^. 

7) From the known values of T and ffiaer > determine the terms Aj and By 
entering the right-hand side of the differential equations (see Table 4«l)» 

8) To determine the value of c,^. it is necessary to know the values of 
blade twist in the feathering hinge Yo ' ^°^ ^-^ ^b blades of the rotor. In 
this case, c, , is determined by the method presented in Subsection 5. 

9) Determine the value and sign of the moment of friction Mfp in the 
feathering hinge (see Subsect.9). 

After this, derive all coefficients of the equations (see Table 4«l) and 
start with the solution. 

10) The system of equations (see Table 4.1) permits determining all values 
of 5j and Ytw i^ ^j » Ytw aJ^d the right-hand sides of the equations Aj and Bjj 

are known at the azimuth ^^ in question. This fact permits its use in the /432 

calculation program in the form of some operator of the type 



h'\=p(h'\„,v- ^^-^^^ 



After applying this operator, determine the values 6^ and y'tw ^ 'the initial 
instant of time. 

11) The change to the next instant of time can be accomplished by means of 
various methods of numerical integration of differential equations. 

468 



Good results are obtained by a system of formulas in which the transition 
from the instant of time t to the time t + At is accomplished by two checks. 
This system of formulas is illustrated for the example of determining the values 
of the variable 6j . The index pertaining to the number of the harmonic is 
omitted for simplicity. 



First check: 



Determination of 5.„: 



Second check: 



8<+4< = 8, + A^8,; 

8/+4/ = /'(8)+4,, y\+M, <^t+iii). 



*/ + ti/ + A( 



(7.35) 



8/+A< = 8<-l-A^8^v; 

8<+4/ = /'(8/+i/, y'Va/, <i'<+A/). 

The values of 6t+Atj S^+^t* ^t+At obtained as a result of recalculation are 
considered final for the instant of time t + At . 

Operations analogous to eq.(7.35) are performed on the coefficients of tor- 
sional deformations. The change-over to the next instant of time is thus accu- 
rately accomplished. 

A simpler method of numerical integration can be proposed. This will be 
presented in greater detail in Vol.11. 

12) The type of problem investigated is important for the sequence of 
calculation. If it is a question of determining the possibility of rotor flutter, 
then the process of numerical integration must be carried out simultaneously for 
all rotor blades and the value of 0,, must be determined at each instant of time. 
The coupling between blades through the swashplate is taken into account by 
calculating the quantity c^ ^ . If the question of investigating flutter is not 
raised and only stresses in the blade are being determined, the problem is great- 
ly simplified. In this case we can introduce into the calculation the assump- 
tion that all blades of the rotor duplicate the motion of the blade in question, 
and the process of numerical integration is performed for only one blade. 

When determining c^ , in this case it is assumed that 7433 

1*69 



YMr)=Vo(C-^). (7.36) 

where 

i|r^°^ = azimuth angle of the blade with the ntunber N = 

/ 9 H N whose motion is determined in the calculation; 

Yo('l'b°^ z^j " coefficient of deformation of the blade with the 

'' number N = 0, not at the azimuth %°^ in question but 



at the azimuth i^i°'' g 



YN^'b ) " coefficient of deformation of the blade with the 

number N, when the blade with N = is at the azi- 
muth ^i°' . 

13) In determining the stresses, the nimierical integration is performed for 
several rpm of the rotor until all values of 6^ and Ytw^'t 'two successively 
calculated rpm differ less than the prescribed accuracy of calculation. This 
will indicate that the process has converged. After this, the bending stresses 
at each azimuth can be determined by the formula 



=2v 



(/■) 



(7.37) 



where o^^^ are the bending stresses of the blade with respect to a normed mode 
of natural vibrations of the j-th harmonic. 

Further reduction of the obtained data can be performed in any form, de- 
pending on the purpose of the calculation. Usually, the amplitude of the 
stresses is determined and the variation in stresses with respect to azimuth is 
decomposed into harmonics. 

14) In the investigation of flutter, the results can be evaluated after 
studying the entire process of variation in the deformation coefficients during 
several rotor rpm. This is not very convenient in practice since it requires 
considerable graphic work for plotting the dependences 6j = f(i|f) and Ytw ^ f(t)« 
Nevertheless, these drawbacks are compensated by the advantages of this calcula- 
tion method. 

The method presented here involves a large amount of work, but it is known 
from practical experience in design shops that, if modern digital computers are 
used, this method best meets the requirements in designing and perfecting blades 
and permits introducing additional refinements into the results of the calcula- 
tion based on an analysis of the roots of the characteristic equation. 

Section 8. Experimental Investigations of Flutter 
1. Ground Tests for Flutter 

The features of helicopter design permit the performance of flutter analysis 
470 



of the rotor under safe conditions, with the helicopter on the ground. This 
constitutes a distinct advantage of the helicopter over regular aircraft. 

Ground tests for flutter are carried out for different purposes. Often 
these purposes are purely of a research nature. In many cases, it is necessary 
to check or refine - under full-scale conditions - the effect of various para- 
meters on flutter characteristics, to evaluate the peculiarities of the develop- 
ment and cessation of flutter and, finally, to simply refine individual moments 
in the procedure of conducting such tests. 

Nevertheless, in the overwhelming majority of cases these tests are /434 
carried out for inspection purposes. Recently, it has become the rule that each 
experimental helicopter must undergo flutter tests before the start of flight 
tests. The actual margins to the onset of flutter are established in these 
tests. If they prove to be too large, the designer can reduce them, for example, 
by decreasing the weight of the counterbalance in the blade and thus lightening 
it. In the case of insufficient margins, it is necessary to make some design 
modifications and recheck them in tests. 

The finally established flutter margins on an experimental helicopter will 
later serve as criteria for evaluating the characteristics of other helicopters 
of the same design in production at a series-production plant or in actual 
service . 

Usually, in developing a new helicopter it is possible to restrict the test- 
ing to ground tests without the need for additional flight tests. In exceptional 
cases in the past, it had been necessary to also conduct flight tests. As a 
rule, there is no need for these. 

Ground flutter tests are usually carried out in the following manner: 

The helicopter is made fast on a special platform so that the possible oc- 
currence of flutter and consequent failure of a part will not cause the heli- 
copter to roll over. As is known, roll-over of a helicopter will cause the 
blades to strike the ground and almost completely wreck the craft. In some 
cases, there might be casualties. Generally, such does not happen in ground 
tests for flutter, but the experimenter must always be prepared for any eventu- 
ality. 

To begin the tests, the rotor should be revved to the maxim.um rpm at which 
flutter cannot yet occur. Then the rpm is gradually increased. Usually, this 
increase is accomplished in steps of a certain quantity An, so that ng = ni + 
+ An. Here ni is the initial value of the rpm and ng the new value. The quanti- 
ty An is generally taken as about 2% of the operating rpm of the rotor. 

At the new rpm ng, the rotor is held for some time (usually 1-2 min) so 
that vibrations can proceed up to noticeable intensities; if flutter does not 
occur, the rpm is again increased by the quantity -An until flutter does develop. 

Flutter tests are usually greatly simplified if, to cause flutter, it is 
not necessary to create initial disturbances as it is required in hard flutter 
with an excitation threshold (see Sects. 3. 3 and 3«4). Therefore, an attempt 

471 



% 



should be made in the tests to create conditions favorable for the occurrence of 
soft flutter. Such conditions usually are present if a sufficiently large forced 
motion is generated in the feathering hinge. For this, the control lever, and 
along with it the swashplate, are deflected forward as far as possible. Usually, 
this is limited by the fact that the blades begin to strike the supports of the 
vertical overhang guard. 

When a forced motion is created in the feathering hinge, flutter sets in 
earlier with respect to the rotor rpm. Thus, pulling the control stick, in a 
way, is a means of generating flutter. Here, the start of flutter tests is as 
follows: The increase in rotor rpm by An is carried out at neutral position of/435 
the swashplate, after vrfiich the control stick is pulled forward and the regime 
is maintained with the stick deflected. If flutter does not occur, the control 
stick is returned to the neutral position and the rotor rpm is again increased, 
and so on, until flutter occurs. 

Upon the appearance of flutter, if the oscillations build up rapidly, it is 
first necessary to reduce the engine power sharply so as to cause a rapid drop 
in rotor rpm. An additional means of stopping flutter is to return the control 
stick to the neutral position. 

In flutter tests, it is of great importance to achieve the maximum possible 
rotor rpm. To prevent the rpm from being limited by the engine power, the rotor 
is usually lightened meaning that the angle of blade setting is reduced. Experi- 
ments have shown that the overall angle of blade setting has only a slight in- 
fluence on the critical rpm of flutter and thus can be reduced without risk. 
However, one definite limitation does exist. The lower the angle of rotor set- 
ting, the sooner will the blade begin to strike the supports where the control 
stick is deflected. Furthermore, severe lightening of the rotor is unwarranted 
so that the maximimi rpm in the tests is limited not so much by the power as by 
the mechanical strength of the engine. Therefore, the angle of blade setting is 
selected as maximum in the tests but is kept at a value preventing the blades 
from striking the supports when the control stick is deflected while maintaining 
sufficient engine power for maximum possible rpm allowable for mechanical 
strength reasons. 

Flutter tests under ground conditions obviously are possible only if the 
rotor characteristics are such that flutter will take place under these condi- 
tions. On helicopters rated for service, flutter cannot occxor under ground con- 
ditions. Therefore, to conduct ground tests for flutter, the rotor parameters 
must be disturbed somehow. This is usually accomplished in the simplest way by 
disturbing the blade balance, which can be achieved by attaching small weights 
to the trailing edge of the blade. Occasionally, the balance is shifted by coat- 
ing the surface of the blade close to the trailing edge with some kind of ma- 
terial whose weight will shift the blade balance rearward. It is also possible 
to introduce some elastic elements into the control loop. Thus, in conducting 
flutter tests, the rotor parameters must first be changed so as to make occur- 
rence of flutter possible. 

When conducting the tests, it is necessary to provide for the recording of 
various parameters to permit an accurate determination of critical rpm, fre- 
quency, vibration mode, and deflection of the control stick at which flutter 
began. 
472 



Without a determination of these parameters it is impossible to make a suf- 
ficiently accurate evaluation of the flutter margin and to indicate what para- 
meters should be changed to increase this margin. 

When conducting the tests, the onset of flutter is detected by the pilot 
from the disturbance of the blade coning angle and from the increase in fuselage 
vibrations and, in the case of reversible control, also from vibrations of the 
control stick. However, all these signs are sufficiently distinct only after 
the vibration amplitudes reach extremely high values and conduction of the test 
becomes dangerous. Consequently, it is desirable to stop flutter tests earlier, 
before oscillations have time to develop. In this case, the pilot may easily /436 
confuse flutter with the usually present distortion of the blade coning angle. 
This is promoted by vibrations which, as a rule, arise in such tests owing to 
wind and lack of controllability of the rotor. 

In this case, the occurrence of flutter can be judged- only by recordings of 
various factors that are characteristic for vibration. To determine the onset 
of flutter and its parameters, the type of recordings made in the tests is of 
great importance. It has been shown that it is not always easy to determine the 
onset of flutter from a recording of the flapping motion of the blade in the hub 
hinges, since flutter vibrations in these hinges lead only to a distortion of 
the recording of the flapping motion caused by deflection of the control stick. 

This is illustrated in Figs. 4-33 and 4*34 which show a recording of blade 
motion in the flapping hinge (angle p), with the recording of weak flutter shown 
in Pig. 4. 33 and of stronger flutter in Fig.4.34« As follows from Fig. 4. 33, a 
determination of the onset of low-amplitude flutter would be difficult from a 
recording of the angle p. The same is true with respect to recording the hinge 
moment M^. 

The onset of flutter is best reflected in the recording of forces in the 
nonrotating control loops. The recording of forces in the longitudinal control 
Piong is shown in the oscillograms (Figs. 4*33 and 4.34). It is easy to define 
the onset of flutter from these recordings. 

It should be mentioned that Figs. 4*33 and 4.34 show the recordings of anti- 
phase flutter with an order m = 3 for a four-blade rotor. Consequently, the 
frequency of the variable forces in the longitudinal control is governed by the 
relation 

P}o,g ^^Pjlu + f^fl"^ (8.1) 

where 

jOfi„^330 osc/m'in 

Nfiu^ 184 fpm 

When the control stick is deflected from the neutral position by even the 
slightest amount Xp, the vibration frequency of the blade in flutter Pfju can no 
longer be determined from recording the angle 3 (see Figs.4«33 and 4.34), but 
can easily be calculated from eq.(8.l) since the value of the frequency Piong is 

473 



■p- 




n~n3.rpm 

' I I 



Fig. 4.33 Oscillogram of Hinge Moment M^, Blade Flapping Angle P, Position 
of Control Stick Xp, and Forces in Longitudinal Control Piong in 
the Presence of Weak Flutter. 




■Al'^V''! -vi- '" \Neutral position of stl 



n'One rotor revo lut ion 



\^ \j u \J \J' v; \} y T- J y y IX -'^v } V } • I n_ n'^ne roior revo lution 

:^:lr-^~::zr^- ^ ' ! -J. ^^ J^y^U aX attA 7u&.-A7tKAi^AA rys^p;^^^^^^ at^^ttT'^^-A^-A- m a a a a L ;. ..1. . J . A . I — 



^ ■ « *■ t^1»<ma^^ t t I 'l^f I * ff > 



ri 



-1- 



. Onset of flutter. 

I ! n-isirpm , 

! 1 I I I 



n - rss rpm \ 

I I I I 



,n =186 rpm 



Fig. 4.34 Oscillogram of Hinge Moment M^, Blade Flapping Angle p. Position 
of Control Stick Xp, and Forces in Longitudinal Control Piong i^^ 
the Presence of Stronger Flutter. 



5 

-<3 



readily determined from the oscillograms. 

If, after occiirrence of flutter, the control stick is returned to neutral, 
the flapping motion caused by tilting of the swashplate will stop and the only 
motion in the flapping hinge will be that due to flutter (see Fig.4«34)« In 
this case, the frequency of flutter can be determined also from the recording 
of p. 

In the tests whose recordings are shown in Figs. 4*33 and 4*34, flutter was 
caused by an increase in rotor rpm and by deflection of the control stick by an 
amount Xp. 

In the first case (see Fig. 4. 33), the rpm was raised to nfiu = 184 and, as 
soon as weak flutter set in, it was stopped again by decreasing the rotor rpm. 
The position of the control stick Xp had not been changed. 

In the second case (see Fig. 4. 34), the rpm was raised somewhat more, up to 
n = 186, causing stronger flutter to occur. At the start, the control stick was 
retiorned to neutral without a change in rpm; this caused the increase in vi- /438 
bration to stop, after which the rpm was lowered and the flutter disappeared. 

It should be mentioned that the recordings shown in Fig. 4. 33 and 4.34 cor- 
respond to rather weak flutter with a slowly increasing amplitude. Such flutter 
is not always observed; often, the vibration amplitude increases much more 
rapidly and the manipulation of the control stick, described above, becomes im- 
possible. 

As an example of such abruptly developing flutter. Fig. 4-35 shows an oscil- 
logram of blade motion about the flapping hinge for another helicopter with a 
three-blade rotor. To stop flutter on this helicopter it was necessary to reduce 
the rpm as rapidly as possible. 

The flutter vibration mode whose recording is shown in Fig. 4-35, is of the 
in-phase type which means that the collective pitch control is loaded during the 
vibration. This makes the recording of the swashplate slide vibrations, shown 
in Fig. 4. 36, quite interesting. This recording was made with a CV-11 automatic 
recorder. 

The recordings shown in Figs. 4*33 - 4»36 are given only as an example and 
in no way exhaust all possible types of flutter observed on helicopters. These 
types may differ in modes of blade vibration, phase distribution of vibrations 
over the blades (different values of m), frequencies, rate and character (soft 
and hard flutter) of build-up of vibrations, and in numerous other features. 
All these peculiarities must be taken into account in flutter tests and in pro- 
cessing the obtained recordings. 

2. Flutter Tests in Flight 

Flutter tests in flight became necessary when it was found that, during 
mass service of helicopters, there were individual cases of flutter in flight 
when such flutter should not have been possible according to concepts held at 
that time. 

475 



OS 






10^ I 



15-:? 



ZItO 



ri'ZIIO 



2)10 



One revolu tion of rotor 



m-- 

ffwJ 

/rmt 




Fig. 4. 35 Oscillogram of Blade Flapping Motion during Violent Flutter. 




Onstt of flutter 



End of flutter 



Fig. 4. 36 Recording of Forces in Collective Pitch 
Control during Violent Flutter. 






Tests were carried out, which showed that 
the critical flutter rpm in flight is appreci- 
ably lower than in ground tests. 

The relation between the critical rpm in 
flight and on the ground was calculated and 
it became possible to define the characteris- 
tics, checked in ground tests, that were 
needed for prevention of flutter in flight. 
The obtained conclusions can be used in de- 
veloping new helicopter, making flutter tests 
in flight for each type of helicopter wa- 
necessary. It should be borne in mind that 
tests with excitation of flutter in flight 
are extremely dangerous. Such tests should 
be performed only if absolutely necessary and 
should be organized with maximum safety for 
the crew. 

Primarily, before starting the tests the 
researchers should collect data ensuring that 
abrupt development of flutter will not occur 
in flight and that, if it does start, it can 
be stopped again. Such data can be obtained 
in cases in which unscheduled flutter sets in 
during flight tests or diiring service on some 
helicopter of the type in question. This 
occasionally occurs as a consequence of some 
operating error, for example, if the rotor is 
revved to an rpm by far exceeding the permis- 
sible maximum. 

Ground tests can be used as an indirect 
criterion for the degree of abruptness of 
flutter. Experience has shown that the rate 
of build-up of vibration on the ground and /440 
in flight is determined to some extent by the 
overall parameters. Therefore, in some cases 
data of ground tests can be used as basis. 

The only reliable measure for stopping 
flutter in flight is a sharp reduction in 
rpm. Therefore, to ensure definite stopping 
of flutter it is necessary to have a large 
rpm excess in a regime where flutter begins 
in comparison with the minimum rpm at which 
sible. During the tests, the pilot should induce flutter by raising 
sharply reducing the rpm to the minimum possible for 




bO 
•H 
H 

C 
•H 

& 
■P 

bfl 



13 
bO 

•H 
ft 
ft 
cd 

H 

o 



o 



•H 
o 
m 

O 






P3H 



flight is possxbie. uurxng 
the rpm and stop flutter by 
continuation of the flight. 



All considerations referring to recording in ground tests hold also for 
flight tests. However, we should point out one peciiliarity of vibrations during 

477 



flutter in flight, which distinguishes these vibrations from those observed in 
ground tests. 

In ground tests, forced flapping motion in the hinges caused by tilting of 
the swashplate takes place almost exclusively at the frequency of the first har- 
monic of the rotor rpm. In flight, the flapping motion contains also the second 
and higher harmonics. Therefore, blade vibration in flapping flutter usually 
generated at frequencies close to the second harmonic but generally not equal to 
it, will lead to beats between the second harmonic of flapping and flutter vi- 
bration. Therefore, flutter in flight is often perceived as beats. 

As a typical example. Fig. 4.37 shows the recording of flutter in flight in 
a regime where flapping in the axes of the shaft consists almost exclusively of 
the second harmonic (see Fig.4«37b). This is explained by the fact that attach- 
ment of the shaft was selected such that vibrations of the first harmonic are 
eliminated in cruising flight. 

The vibrations during flutter in this regime have well-defined beats (see 
Fig. 4. 37a). 

In all other cases, if the lower critical rpm is disregarded, flutter in /441 

flight will not differ from that ob- 
served on the ground. 



Iflu fprnj Pflu "^'^In'ln 



300 



too 



too 




No flutter up to maximum 
ro tor rpm, incl . 



23W 



21'/o 



25% 



26 fo 



*W 



Fig. 4. 38 Comparison of Experimental 
and Calculated Values of Vibration 
Frequency and Critical Flutter Rpm. 



3. Comparison of Calculation and Experi- 
ment under Conditions of Axial Flow 
past the Rotor 

In comparing calculation and ex- 
periment, the initial rotor parameters 
used in the calculation are of prime 
importance, along with type of blade 
balancing, rigidity of the control 
system, and magnitude of friction in 
the feathering hinge of the hub, as 
well as reliability with which the lo- 
cation of the profile focus is known. 
Errors in determining the initial data 
naturally affect the accuracy of deter- 
mining the flutter parameters. There- 
fore, in comparing calculation and ex- 
periment it is desirable to eliminate 
errors in determining the initial para- 
meters. For this, the parameters 
should be checked experimentally. 



Balancing should be determined by weighing individual segments of the blade ob- 
tained after cutting it. 

To determine the control rigidity a special method of measuring dynamic 
rigidity should be used, which will be taken up in greater detail in Section 6. 
The use of other methods generally leads to misunderstandings and fallacies and 

478 



therefore should be discarded. 



To check the position of the profile a.c. a segment of a full-scale blade 
should be exposed to the air stream in a wind tunnel. In this case, it can be 
expected that deviations in the aerodynamic characteristics due to design errors 
of the blade profile and deformation in work will be refined to some extent. 

Figure 4*38 gives the results of a comparison of calculation and experiment 
for the Mi-4 helicopter. The solid curve shows the theoretically obtained de- 
pendence of the critical flutter rpm on the effective blade balancing. The 
circles mark the experimental results. Circle 1 with the forwardmost blade 
balancing corresponds to the maximum rpm obtainable with a helicopter engine. 
There was no flutter in this case. After attaching 0.46-kg weights to the blade 
flaps, the experiment was repeated. There again was no flutter (circle 2). 

Attachment of weights of 0.86 and 1.3 kg to the blade flap caused flutter 
at rotor rpm of n = 18? and n = 173 respectively (squares 3 and 4 in Fig. 4. 38). 

The frequency of blade vibration during flutter is indicated in the dia- /442 
gram by squares 5 and 6, which should be compared with the theoretically deter- 
mined frequency values shown by the dashed 
curve. 

After the experiments, the blades were 
cut into segments and their effective 
balancing was determined, which is noted on 
the graph in Fig. 4. 38. The dynamic rigidity 
of the control system was determined on the 
same helicopter. The magnitude of friction 
in the feathering hinge, which was highly 
stable, was measured in the laboratory on 
another hub of the same design. 

These data indicate satisfactory (with 
an accuracy to within 0.^% of the chord for 
the value of effective balancing) agreement 
of calculation and experiment. We note that 
such a good agreement was observed in all 
other experiments carried out on other heli- 
copters. This creates confidence in the 
reliability of the results obtained from 
calculation and in the validity of the 
initial assumptions, including that of the 
permissibility of determining aerodynamic 
forces by formulas based on the "steady- 
state hypothesis" . 



"flu rpm 














■////// 






/ 
/ 


/ 






uoa 




/ 
/ 


/ 


/ 






1 

1 
i 
1 

- V 

\ 


/ 


/ 








300 


/I 
















200 


>W/Vj 


■////J 


"m 
/////, 


ax 




s 


hift 


of 2 


II V 


>>^ 


t 




=0- 


100 


bal ancing 

due to a 

- n.86-kg weig 


j 




'""■-- 


^ 






















f^ 


u,cj 





















23% 



21% 



25% 



26% 



'«// 



Fig. 4. 39 Comparison of Experi- 
mental and Calculated Data in 
Flight. 



It should be added that the flutter calculation pertains to a case quite 
rare in rotor calculations when good agreement vd-th experiment is observed. 
Probably, this is due primarily to the fact that even substantial errors in de- 
termining the magnitudes of aerodynamic forces have no great effect on the final 
results of calculation at critical flutter rpm. 

479 



4. Comparison of Calcula tion and Experiment in F light 

Comparisons of calculation and experiment in flight do not show such good 
agreement as in similar comparisons of results obtained under conditions of 
axial flow past the rotor in ground tests. In flight, the decrease in critical 
flutter rpm is felt more strongly than on the basis of calculation. Figure 4«39 
gives two curves obtained by calculation for a regime with axial flow (ij, = O) 
and for horizontal flight with p. = 0.25. The curves do not differ greatly. 
Conversely, the experimental results differ substantially. In Fig. 4. 39 point 1 
marks the critical rpm obtained in a ground test with a 0.86-kg weight attached 
to the flaps while point 2 refers to the critical flutter rpm obtained on the 
same helicopter in flight but without weights on the flaps. The test points in 
Fig. 4. 39 were obtained in tests laid out by S.B.Bren and A.A.Dokuchayev and per- 
formed by the pilot V.V.Vinitskiy. 

The diagram indicates that the difference between the flight and ground 
tests is appreciably greater than that obtained by calculation. The cause for 
the difference lies in the fact that, in calculations, the amplitude of the /443 
forced motion in the feathering hinge was taken to be the same on the ground and 
in flight, i.e., it was assumed that in ground tests the amplitude of the angular 
velocities of blade vibration in the feathering hinge, as a result of deflecting 
the control stick, was the same as in flight as a consequence of ordinary flap- 
ping motion. Here, it was disregarded that, in flight, the different vibrations 
and oscillations with harmonics of higher orders may noticeably reduce the ef- 
fectiveness of damping of oscillations due to friction in the feathering hinge. 
This assumption is usually made to explain the more abrupt drop in critical 
flutter rpm in flight in comparison with the calculation. 

5. Check for Flutter 

It has been noted above that, for a reliable elimination of the possibility 
of flutter under service conditions, the helicopter rotor should have a well- 
defined flutter margin. This margin should be checked on the ground and, if the 
margin is below some standard value, the helicopter should not be allowed to fly. 
In this approach, the required margin before flutter, checked on the ground, 
should take into account a decrease in critical rpm in flight, possible deterio- 
ration in flutter characteristics due to moisture penetrating into the blade, 
and other factors, and should secure the necessary stability of blade vibration 
at maximum approach to this margin. 

The idea of flutter checking was first expressed by M.L.Mil' who proposed 
to excite rotor oscillations by installing an eccentric in the cyclic pitch con- 
trol system and to measxire the stability margin in terms of the amplitude of the 
obtained resonance vibrations, which should be greater the smaller the flutter 
margin. Such experiments were carried out and yielded interesting results. 

Figure 4.40 shows the experimentally obtained dependence of the amplitude 
of the hinge moment on the excitation frequency of the eccentric for various 
rotor rpm. The diagram shows that the' higher the rotor rpm and hence the closer 
to flutter, the greater will be the amplitude of the hinge moment. The same de- 
pendence is obtained for blade balancing. Dm^ing experiments on the Mi-4 heli- 

480 



HO 

n^l78rpm 



copter vri-th a four-blade rotor, we noted the occurrence of two modes of resonance 
vibrations of frequencies Pi = Pgco + n and Ps = Pecc - n (pgoc is the frequency 
of excitation from the eccentric^, which agrees nicely with the theoretical 
notions presented in Section 4» The experiments confirm the possibility of using 
the described method for checking the stability margin of the rotor. 

However, some time later a simpler method for checking the necessary margin 
in terms of blade balancing was developed. This method provides for checking 
the helicopter on the ground with blades whose balancing is shifted rearward by 

a certain predetermined quantity. 
The balance is shifted by means 
of special weights placed on the 
trailing edges of the blade 
during the check. If, on raising 
the rpm to a prescribed maximum, 
flutter does not set in, the 
weights are removed and the heli- 
copter is admitted to service. 

The weights were originally 
selected on the basis of /hhh 
calculations and later corrected 
for different experiments and 
service conditions. Two magni- 
tudes of the required margins are 
usually established. When the 
helicopter is released from the 
plant, an increased margin is 
established -vdiich can be partial- 
ly expended in service. There- 
fore, in a number of cases flutter 
check is also introduced in 
service, but then a smaller re- 




300 too p^^osc/min 



Fig. 4. 40 Hinge Moment Amplitude as a 
Function of Vibration Frequency of the 
Eccentric. 



quired margin is established. 

The introduction of a flutter check has proved a useful measure, after 
which cases of the development of flutter in service completely stopped. 



6. Experimental Determination of Control System Rigidity 

It was already pointed out above that the critical flutter rpm greatly de- 
pends on the magnitude of the control system rigidity. It can be approximately 
assumed that the critical flutter rpm is directly proportional to jCaon • Hence 
it is obvious that it is important to determine control rigidity as accurately 
as possible for a successful calculation. How does one determine the magnitude 
of this rigidity? When performing the first calculations for flutter, control 
rigidity is often calculated theoretically by stmming the design rigidities of 
all components entering the control loop. First measurements of this rigidity 
showed that the calculated values are much higher than the experimental values. 
Therefore, it was necessary to reject the calculation of control rigidity. 



481 



However, the problem of the manner of experimental determination of control 
rigidity also proved difficult. At first, the control rigidity was determined 
statically, i.e., by the slope of the dependence of the magnitude of deformations 
on the external load. However, this method did not clarify the mode of account- 
ing for play in the control system, friction, and inertia of the components 
entering this system. Therefore, the so-called dynamic method of determining 
control rigidity was used, in which the external forces exerted at the control 
by the blades were applied dynamically, at a frequency equal or close to the 
frequency of flutter. With this method of measurement, the control rigidity /kh5 
was by a factor of 2 - 2.5 less than with the static method. 

It is natural that the results obtained in static analysis cannot be used 
for the flutter calculations. 

What is the simplest way of determining the dynamic rigidity of the control 
system? For this, we used the following method: 

On a helicopter with a nonrotating rotor we replaced the blades by special 
weights whose moments of inertia relative to the feathering hinge were equal to 
the moments of inertia of the removed blades. By measuring the natural vibra- 
tion frequency of this system, the magnitude of the corresponding hinge control 
rigidity can be completely defined. These rigidities can obviously be calculated 
by means of the formula 



where 



Ccon-P'h^p, (8.2) 

p = one of the natiiral vibration frequencies of this system, which 

should be considered equivalent to a rotor with blades absolutely 
rigid in torsion; 
Coon ~ hinge control rigidity corresponding to the vibration mode for 
which the frequency p is determined. 

The necessary values of the natural vibration frequencies can be determined 
by the usual method of forced vibrations with excitation by a vibrator or ec- 
centric . 

Since the rigidity of the longitudinal and lateral controls on a helicopter 
is usually not the same, two different values of the natural vibration frequency 
will correspond to loading of these controls [see eq.(4.19)]. 

Let us present the values of the frequencies corresponding to loading of 
different control loops obtained on the Mi -4 helicopter with a nonrotating rotor: 

/»^=400^420 osc/mi» 

/7j=440-^450 osc/min 

p^^ =590-1- 620 «wc/«/rt 

p =920-^940<?J?<;/w/n. 



482 



The notations used here are the same as those used in eqs.(4«19)» 

This raises the question whether the control rigidity thus measured depends 
on the amplitude of external forces acting on the control system. To check this, 
we carried out experiments with the maximum pennissible (in terms of strength) 
magnitudes of hinge moments acting on the control, approximately the same as 
those which act at the maximum flying speed, and witH moments lower by a factor 
of 10. There was no substantial difference in the value of the obtained fre- 
quencies. 



Dynamic control rigidity may depend on the frequency of the forces acting 
in the control cables. By changing the moments of inertia of the weights in- 
stalled in place of the blades and measuring the new natural vibration frequen- 
cies of the system, it becomes possible to define the mode of variation of 
rigidity with variation of the vibration frequency. Figure 4.41 shows the re- 
sults of such measurements. The abscissa gives the natural vibration frequency 

for the control system which varies as a 
function of the magnitude of the moment of 

Px'P^x-pe^f osc/min^ ^ ^ ^ ^ inertia of the weights, while the ordinate 

gives the dynamic rigidity expressed in 
terms of the corresponding natural vibra- 
tion frequency in agreement with eqs.(4.19). 

Figure 4.41 indicates the approximate 
values of the frequencies of variable /446 
forces acting' in the nonrotating parts of 
the lateral and longitudinal controls 
during flutter with modes of the first 
(m = l) and third order (m = 3). These 
results illustrate the above assumption 
(see Sect .4.3) that the magnitude of con- 
trol rigidity may depend on the frequency 
200 100 too posc/m/n of the forces acting in it. 




200 



Fig. 4. 41 Control Rigidity as a 
Function of Vibration Frequency. 



of effective forces. 



For comparison. Fig. 4. 41 also gives 
the values of the static control rigidity 
obtained from the slope of the dependence 
of control deformations on the magnitude 



The dynamic method described here for determining control rigidity has been 
sufficiently checked and can be recommended for practical use. 

7. Experiments on Dynamically Similar Models 

For conducting experiments on full-scale helicopters, the researcher usual- 
ly runs into many difficulties having to do with observance of safety rules, 
since full-scale experiments are usually carried out by a pilot or mechanic in 
the helicopter. This imposes certain restrictions, especially for flutter tests 
in flight where, for safety considerations, flutter is usually generated only 
once in some regime or, in the extreme case, three to four times but never more 



4^3 



often. It is impossible to obtain any dependences for the parameter. 

Furthermore, there are limitations to the possibility of investigating 
various flight regimes, due to the characteristics of the helicopter on which 
the experiment is carried out. The engineer is almost always interested in the 
flutter margin with respect to rpm. However, the maximum rpa achievable in ex- 
periments is limited by the capabilities of the engine. For example, the maxi- 
mum flying speed is limited. Therefore, the researcher naturally will attempt 
to make wind-tunnel tests on dynamically similar models. Such tests often jrield 
interesting results. However, their wide use is restricted by a number of basic 
shortcomings. To estimate the need for such tests in each individual case, let 
us discuss the basic principles underlying the simulation in greater detail. 

In producing a reduced-scale rotor model, geometric similitude of the ex- 
ternal blade shape and the characteristic linear dimensions of rotor blade and 
hub are of prime importance. We are thinking here of linear dimensions deter- 
mining the planform of the blade; distribution of profiles and their setting 
angles over the blade length; dimensions of its components determining, for 
example, the position of the feathering hinge axis along the blade length; /447 
relative position of other hub hinges; and many other dimensions. Next, it is 
necessary that all relations between aerodynamic, inertia, and elastic forces 
remain constant. In this case, the variable aerodynamic loads set up at the 
model blade lead to the same relative deformations as on the original blade. 

Let us examine this in greater detail for the example of bending vibrations 
of a blade in the flapping plane. It can be demonstrated that bending deforma- 
tions of a blade with respect to some natioral vibration harmonic are determined 
by the coefficients of deformation calculated by the formula (see Vol. II) 

Pj 

where 

pj = frequency of the j-th harmonic of natural blade bending vibration; 
Yj = mass characteristic of the blade in vibrations of the j-th har- 
monic [see eq.(7.55) of Chapt.I in Vol.11] 

-^'^;e*o.7/?' 
Y;=-^— ; (8.4) 

mj 

ck'j = dimensionless coefficient, characterizing the magnitude of work 

done by the aerodynamic forces in displacements of the blade during 
deformation with respect to the j-th harmonic: 



a^ = JPt/<^'afr. (8.5) 



Let us define the mode of variation in the relative coefficients of blade 
bending deformations 6^^^ upon a similar change in all its geometric dimensions. 

484 



J 



The relation between aerodynamic and inertia parameters of the blade is de- 
termined by the values of the mass characteristics of the blade Yj • If all geo- 
metric dimensions of the blade change the same number of times, namely, Kl times, 
then, as follows from eq.(8.4), the mass characteristics of the blade do not 
change. 

However, we see from eq.(8.3) that, to retain similitude in bending de- 
formations, the relation between the natural vibration frequency pj and the 
angular velocity of rotation of the rotor cu must be retained. This requirement 
is equivalent to keeping the Strouhal number constant: 

where 

p = vibration frequency; 
U = velocity of flow. 

The natural vibration frequency pj is determined by the formula 



n2- 1 



P-=- 



lEl{yJdr-^^N(yJdr 



(8.7) 



Upon a similar change in all geometric dimensions of the blade, the quantity 
of the elastic moment of inertia of its section I changes K* times. In this 
case, as easily seen from eq.(8.7)> the magnitude of the natural vibration fre- 
quency of the nonrotating blade po, changes Kl times. Consequently, the rela- 
tion between this frequency and the angular velocity of rotation remains con- /kiS 
stant if the angular velocity changes the same number of times. 

Thus, to retain similitude in aerodynamic, inertia, and elastic forces, all 
geometric blade dimensions must change the same number of times (Kl) and the 
peripheral blade speeds must remain constant. Such dynamically similar models 
are called Mach-similar models since similarity with respect to the Mach number 
is retained in all blade sections. 

The requirement of changing all geometric dimensions the same number of 
times is easiest to meet by keeping the blade design unchanged. Therefore, the 
development of such models actually reduces to the development of models similar 
in design. This is a difficult problem, requiring the solution of many highly 
complex technical problems and the organization of a special production of small- 
dimension designs. A sufficiently high accuracy is necessary in their manu- 
facture. Considerable difficulties also arise in developing hub hinges. It is 
necessary to state that such models are also under considerable stress relative 
to mechanical strength and do not permit much widening of the regimes in which 
investigations can be carried out, in comparison with those on full-scale heli- 
copters. 

Upon a reduction of the geometric blade dimensions, the relation between 
the blade weight and its aerodynamic and elastic characteristics drops by Ki. 
times. This leads to a reduction of the influence of the blade weight para- 
meters in comparison with the value for a full-scale helicopter. In particular, 

IS5 



the relative overhang of the blade of a nonrotating rotor decreases hy Ki times. 
The blade, so to speak, becomes more rigid "to the eye". However, this dis- 
turbance of similitude is observed only when the rotor is not rotating. Upon 
rotation of the rotor the effect of the weight forces is generally negligible. 
Therefore, a disturbance of their similitude has practically no effect on the 
behavior of the blades. 

The difficulties in developing Mach-similar and design-similar blades re- 
sulted in their being used infrequently. Most often, dynamically similar blades 
are developed with disturbance in similitude relative to the Mach ntm±ier. The 
peripheral blade speeds on a model are reduced in comparison with the full- 
scale blade by several times. In so doing, to retain the ratio of natural blade 
vibration frequency pj to angular velocity of rotation cu, the blade rigidities 
are reduced not by Kt times, as is required by geometric similitude, but by a 
greater nimiber of times, most often by Kl. In this case, the necessary ratio of 
natural vibration pj to angular velocity is achieved at peripheral speeds ^K^ 
smaller than those on a full-scale helicopter. Presumably, the results of tests 
on such models can be extrapolated in totality to full-scale units only at 
M < 0.4 (see Fig. 4*3) • At M = 0.5 - 0,9, the test results of such models can be 
used only for qualitative estimates. In this connection, non-Mach-similar models 
are used in only a limited volume for practical purposes. 



486 



REFEEtMCES /AA9 

1. Zhukovskiy, N.Ye.: Theoretical Determination of the Effect of Wind Blowing 

in the Plane of a Helicopter Rotor (1909) [Opyt teoreticheskogo opredele- 
niya effekta vetra, dujrushchego v ploskosti gelikopternogo vinta (1909) ]• 
Collected Works (Sobr. soch.), Vol.VI, p. 68, 1937; Vol. IF, p. 388, 1949. 

2. Yur^yev, B.N.: Aeroldnamic Calculation of Helicopters (Aerodinamicheskiy 

raschet vertoletov) . Oborongiz, 1956 . 

3. Yur»yev, B.N.: Investigation of the Flight Characteristics of Helicopters 

(issledovaniye letnykh svoystv gelikopterov) . Tr. WIA im. Zhukovskogo, 

No. 49, 1939. 

4. Mil», M.L.: Takeoff Run of an Autogiro (O razbege avtozhira). Tekhnika 

Vozdushnogo Flota, No. 5, 1934« 

5. Mil', M.L.: Autogiro Balancing and Stability (Balansirovka i ustoychivost' 

avtozhira). Tekhnika Vozdushnogo Flota, No. 10, 1934. 

6. Mil', M.L.: Dynamic Twisting of an Autogiro Rotor Blade in Flight (0 dina- 

micheskom zakruchivanii lopasti rotora avtozhira v polete). Tekhnika 
Vozdushnogo Flota, No. 2, 1937. 

7. Mil', M.L.: The Aerodynamics of an Autogiro Rotor in Curvilinear Flight 

(Aerodinamika rotora avtozhira pri krivolin^nom dvizhenii). Tekhnika 
Vozdushnogo Flota, No.l, 1939. 

8. Mil', M.L.: Aerodynamics of a Rotor with a Hinged Blade in Curvilinear 

Flight (Aerodinamika nesushchego vinta s sharnirnym krepleniyem lopastey 
pri krivolineynom dvizhenii). Tr. Tsentr. Aero-Gidrodinam. Inst., No .465, 
1940. 

9. Mil', M.L. and Yaroshenko, V.N.: Aerodynamic Calculation of a Helicopter 

(Aerodinamicheskiy raschet gelikoptera) . Tekhnika Vozdushnogo Flota, 
No. 11, 1946. 

10. Mil', M.L.: Helicopters (Vertolety). Izd. Znaniye, 1957. 

11. Bratiikhin, I.P.: Autogiros (Avtozhiry). Gosmashmetizdat, 1934* 

12. Bratukhin, I. P.: Design and Structure of Helicopters (Pixiyektirovaniye i 

konstiTjktsiya vertoletov) . Oborongiz, 1956. 

13. Proskuryakov, A. P.: Effect of Unsteady Flow on the Aerodynamics of the 

Autogiro Blade (Vliyaniye nestatsionarnosti potoka na aerodinamiky lopasti 
avtozhira). Tr. Tsentr. Aero-Gidrodinam. Inst., N0.46O, 1939. 

14. Proskuryakov, A. P.: Aerodynamic Calculation of a Rotor with a Blade Pitch 

Variable in Azimuth (Aerodinamicheskiy raschet nesushchego vinta s pere- 
mennym po azirautu uglom ustanovki lopastey). Tr. HI, No. 16, 1946. 

15. Mikhaylov, A.N.: Theory of the Autogiro Rotor with a Variable Setting 

Angle (Teoriya rotora avtozhira s peremennym uglom ustanovki). Tekhnika 
Vozdushnogo Flota, No .3, 1940. 

16. Baskin, V.E. : Induced Velocities of a Propeller Exposed to Flow at an Angle 

to its Axis (Indiiktivnyye skorosti vozdushnogo vinta, obduvayemogo pod 
uglom k yego osi). Report at the All-Union Congress on Theoretical and 
Applied Mechanics, Moscow, I960 (Doklad na Vsesoyuznom s"yezde po teore- 
ticheskoy i prikladnoy mekhanike, Moskva, I960) . 

17. Baskin, V.E., D'yacheriko, A.S., Maykopar, G.I., and Martynov, A.I.: Inves- 

tigation of Airflow and Loads on the Helicopter Rotor Blade in Horizontal 
Flight (issledovaniye techeniya vozdukha i nagruzok lopasti vinta verto- 

487 



leta V gorizontal^nom polete). Inzh, Zh., Vol. Ill, No. 3, Akad. Mauk SSSR, 

1963. 

18. Maykopar, G.I.: Application of the Vortex Rotor Theory (Prilozheniye vikhr- 

revoy teorii vinta). Tr. Tsentr. Aero-Gidrodinam. Inst., No. 613, 1947. 

19. Maykopar, G.I.: Vortex Rotor Theory (Vikhrevaya teoriya nesushchego vinta). 

Collection of Worics on Propeller Theory (Sbornik rabot po teorii vozdusbr 
rykhvintov). BNI Tsentr. Aero-Gidrodinam. Inst., 1958. 

20. Vil»dgrube, L.S.: Determination of the FlLght Characteristics of a Heli- 

copter in all Regimes and at all Altitudes and Selection of its Design 
Parameters (Opredeleniye letnykh kharakteristik gelikoptera na vsekh 
rezhimakh i vysotakh i vybor yego parametrov pri proyektirovanii) . Tekhr- 
nika Vozdushnogo Flota, No. 6, 1947* 

21. Braverman, A.S.: Theory of an Ideal Helicopter Rotor (Teoriya ideal»nogo 

nesushchego vinta vertoleta). Izv. Akad. Nauk SSSR, Otd. Tekhn. Nauk, 
Mekhan. i Mashinostr., No. 2, 1959. 

22. Wan Shi-Tsun: Generalized Vortex Rotor Theoiy. Problems of Helicopter 

Rotor Aerodynamics (Obobshchennaya vikhrevaya teoriya nesushchego vinta. 
Voprosy aerodinamiki nesushchego vinta vertoletov). Oborxjngiz, 1961. 

23. Durend, V.F.: Aerodynamics (Aerodinamika) . Vol. IV, Oborongiz, 1939. 

24. Hessow and Meyers: Aerodynamics of the Helicopter (Aerodinamika vertoleta). 

Oborongiz, 1954* /450 

25. Peyn, P.R. : Dynamics and Aerodynamics of the Helicopter (Dinamika i aero- 

dinamika vertoleta) . Oborongiz, 1963. 

26. Tinyakov, G.A.: Piloting the Helicopter (Pilotirovaniye vertoleta). Voye- 

nizdat, I960. 

27. Izakson, A.M.: Soviet Helicopter Engineering (Sovetskoye vertoletostroyeni- 

ye). Oborongiz, 1954« 

28. Golubev, V.V.: Lecture on Airfoil Theory (Lektsii po teorii kiyla) . Izd. 

Tekhn.-Teoret. Lit., 1949- 

29. Keldysh, M.V.: Hydrodynamic Derivation of Rauscher^s Formula (Gidrodina- 

micheskiy vyvod formuly Raushera). Tekhn. Zametki, No. 52, Part 2, Tsentr. 
Aero-Gidrodinam. Inst., 1935. 

30. Nekrasov, A.I.: Airfoil Theory in Unsteady Flow (Teoriya kryla v nestatsi- 

onarnom potoke). Akad. Nauk SSSR, 1947. 

31. Lamb, G. : Hydrodynamics (Gidro dinamika). OGIZ-Gostekhizdat, 1947. 

32. Grossman, Te.P.: A Course on Vibrations of Aircraft Components (Kurs vib- 

ratsiy chastey samoleta) . Oborongiz, 1940. 

33. Riz, P.M. and Pozhalostin, A.I.: Vibrations and Dynamic Strength of Pro- 

pellers (Vibratsii i dinamicheskaya prochnost* vozdushnykh vintov). Tr. 
Tsentr. Aero-Gidrodinam. Inst., No. 609, 1947. 

34. Ryzhik, I.M. and Gradshteyn, M.S.: Tables of Integrals, Schemes, Series, 

and Products (Tablitsy integralov, skhem, ryadov i proizvedeniy) . GITTL, 

1951. 

35. Watson, G.N.: Theory of Bessel Functions. Izd. Inostr. Lit., 1949- 

36. Glauert, H.A.: General Theory of the Autogiro. R & M, No. 1111, Nov. 1926. 

37. Lock: Further Development of the Autogiro Theory. R & M, No. 1127. 

38. Shapiro, Jacob: Principles of Helicopter Engineering. Temple Press Ltd., 

Bowling Green Lane, London, E.C.I. 

39. Coleman, R.P.: A Preliminary Theoretical Study of Helicopter Blade Flutter 

Involving Dependence upon Coning Angle and Pitch Setting. NACA Rept., 
No .408, 1946. 

40. Miller, R.H. and Ellis, C.W.: Blade Vibration and Flutter. J. Am. Heli- 

488 



copter Soc, Vol.1, pp.l9-3S, JuHy 1956. 

41. Daughaday, H., DuWaldt, F., and Gates, .C: Investigation of Helicopter 

Blade Flutter and Load Amplification Problems. Inst. Aeron. Sci., Pre- 
print, No. 705, Jan. 1957; J« Am. Helicopter Soc, Vol.2, No.3, Jw 1957- 

42. Hooper, M.S.: The Theoretical Approach to some Rotor Blade Flutter and 

Forced Vibrations Problems. J. Helicopter Assoc. Great Britain, Vol.12, 
1958. 

Translated for the National Aeronautics and Space Administration by the 
O.W.Leibiger Research Laboratories, Inc. 



F-49'+ NASA-Langley, 1967 2 489 



ft 



"Tke aeronautical and space activities of the United States shall be 
conducted so as to contribute . . . to the expansion of human knowl- 
edge of phenomena in the atmosphere and space. The Administration 
shall provide for the widest practicable and appropriate dissemination 
of information concerning its activities and the results thereof." 

— Nauonal Abronauxics and Space Act of 1958 



NASA SCIENTIFIC AND TECHNICAL PUBLICATIONS 



TECHNICAL REPORTS: Scientific and technical information considered 
important, complete, and a lasting contribution to existing knowledge. 

TECHNICAL NOTES: Information less broad in scope but nevertheless of 
importance as a contribution to existing knowledge. 

TECHNICAL MEMORANDUMS: Information receiving limited distribu- 
tion because of preliminary data, security classification, or other reasons. 

CONTRACTOR REPORTS: Scientific and technical information generated 
under a NASA contract or grant and considered an important contribution to 
existing knowledge. 

TECHNICAL TRANSLATIONS: Information published in a foreign 
language considered to merit NASA distribution in English. 

SPECIAL PUBLICATIONS: Information derived from or of value to NASA 
activities. Publications include conference proceedings, monographs, data 
compilations, handbooks, sourcebooks, and special bibliographies. 

TECHNOLOGY UTILIZATION PUBLICATIONS: Information on tech- 
nology used by NASA that may be of particular interest in commercial and other 
non-aerospace applications. Publications include Tech Briefs, Technology 
Utilization Reports and Notes, and Technology Surveys. 



Details on the availability of these publications may be obtained from: 

SCIENTIFIC AND TECHNICAL INFORMATION DIVISION 
NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 

Washington, D.C. 20546